Download as pdf or txt
Download as pdf or txt
You are on page 1of 222

Design, Integration Schemes, and Optimization of

Conventional and Pressurized Oxy-coal Power


Generation Processes
by
Hussam Zebian
S.M., Massachusetts Institute of Technology (2011)
Submitted to the Department of Mechanical Engineering
in partial fulfillment of the requirements for the degree of ARCHNE*
MASSACHUSETTS INST"UrE
Doctor of Philosophy OF TECHNOLOGY

at the MAY 0 8 2014


MASSACHUSETTS INSTITUTE OF TECHNOLOGY LIBRARIES
February 2014

© Massachusetts Institute of Technology 2014. All rights reserved.

A uth or .... ..........................................................


Department of Mechanical Engineering
November 1, 2013

Certified by.......
Alexander Mitsos
Visiting Scientist
7') Thesis SUpe;visor

Accepted by .
David E. Hardt
Chairman, Department Committee on Graduate Thesis
2
Design, Integration Schemes, and Optimization of
Conventional and Pressurized Oxy-coal Power Generation
Processes
by
Hussam Zebian

Submitted to the Department of Mechanical Engineering


on November 1, 2013, in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy

Abstract
Efficient and clean electricity generation is a major challenge for today's world. Multi-
variable optimization is shown to be essential in unveiling the true potential and the
high efficiency of pressurized oxy-coal combustion with carbon capture and sequestra-
tion for a zero emissions power plant (Zebian and Mitsos 2011). Besides the increase
in efficiency, optimization with realistic operating conditions and specifications also
shows a decrease in the capital cost.
Elaborating on the concept of increasing the performance of the process and the
power generation efficiency, as part of this Ph.D. thesis, new criteria for the opti-
mum operation of regenerative Rankine cycles, are presented; these criteria govern
the operation of closed and open feedwater heaters, and are proven (partly analyti-
cally and partly numerically) to result in more efficient cycle than the conventional
rules of thumb currently practiced in designing and operating Rankine cycles. Simply
said, the pressure and massflowrate of the bleed streams must be selected in a way
to have equal pinch temperatures in the feedwater heaters. The criteria are readily
applicable to existing and new power plants, with no associated costs or retrofitting
requirements, contributing in significant efficiency increase and major economical and
environmental advantages. A case study shows an efficiency increase of 0.4 percentage
points without capital cost increase compared to a standard design; such an efficiency
increase corresponds to an order of $40 billion in annual savings if applied to all Rank-
ine cycles worldwide. The developed criteria allow for more reliable and trustworthy
optimization, thus, four additional aspects of clean power generation from coal are
investigated.
First, design and optimization of pressurized oxy-coal combustion at the systems-
level is performed while utilizing a direct contact separation column (DCSC) instead
of a surface heat exchanger for more reliable and durable thermal recovery. Despite
the lower effectiveness compared to a surface heat exchanger, optimization employing
newly developed optimal operating criteria that govern the DCSC allow for an efficient

3
operation, 3.8 percentage points higher than the basecase operation; the efficiency of
the process utilizing a DCSC is smaller than that utilizing a surface heat exchanger
but only by 0.32 percentage points after optimization. Optimization also shows a
reduction in capital costs by process intensification and by not requiring the first flue
gas compressor in the carbon sequestration unit.
Second, in order to eliminate performance and economical risks that arise due to
uncertainties in the conditions that a power generation process may be subjected to,
the designs and operations that allow maximum overall performance of the process
while facing all possible changes in operating condition are investigated. Therefore,
optimization under uncertainty in coal type, ranging from Venezuelan and Indonesian
coals to a lower grade south African Douglas Premium and Kleinkopje coal, and in
ambient conditions, up to 10'C difference in the temperature of the cooling water,
of the pressurized oxy-coal combustion are performed. Using hierarchic optimization
and stochastic programing, the latter shown to be unnecessary, an ideally flexible
design is attained, whereby the maximum possible performance of the process with
any set of input parameters is attained by a single design. While in general a process
designed for a specific coal has a low performance when the utilized coal is changed,
for the pressurized oxy-coal combustion process presented herein, it is demonstrated
that designing (and optimizing) while taking into consideration the different coal
types utilized, results for each coal in performance that is equal to the maximum
performance obtained by a design dedicated to that coal.
The third aspect considered is flexibility with respect to load variation. Particu-
larly with the increase of the power generation from intermittent renewable energy
sources, coal power plants should operate at loads far from nominal, down to 35%.
In general this results in efficiency significantly lower than the optimum. Therefore,
while keeping the turbine expansion line design fixed to that of the nominal load in
order to allow for a full range of thermal load operations, an elaborate study of the
variations in thermal load for pressurized oxy-coal combustion is performed. Here
too optimization of design and operation taking into consideration that load is not
fixed results in a process that is flexible to the thermal load; the range of thermal
load considered is 30..100%.
The fourth aspect considered is a novel design for heat recovery steam generator
(HRSG), which is an essential part of coal power plants, particularly oxy-coal com-
bustion. It is the site of high temperature thermal energy transfer, and is shown
to have potential for significant improvements in its design and operation. A new
design and operation of the HRSG that allow for simultaneous reduction in the area
and the flow losses is proposed: the hot combustion gas is splitted prior to entering
the HRSG and prior to dilution with the recycling flue gas to control its tempera-
ture as dictated by the HRSG maximum allowed temperature. The main combustion
gas flow proceeds to the HRSG inlet and requires smaller amounts of dilution and
recycling power requirements compared to the conventional no splitting operation.
The splitted fraction is introduced downstream at an intermediate location in the
HRSG; the introduction of the splitted gas results in increasing the temperature of
the flue gas and the temperature difference between the hot and the cold streams
of the HRSG, particularly avoiding small temperature differences which require the

4
most heat transfer area. Results include area reduction by 37% without change in
the compensation power requirements, or a decrease in the compensation power re-
quirements by 18% (corresponding to 0.15 percent points of the cycle efficiency) while
simultaneously reducing the area by 12%.

Thesis Supervisor: Alexander Mitsos


Title: Visiting Scientist

5
6
Acknowledgments

First and foremost I would like to thank my advisor Professor Mitsos for all his
efforts and contributions in making this work possible. As an advisor, he is attentive,
critical, and extremely generous with his time and thoughts. He always welcomes

and encourages new thoughts and ideas, redefining the principles of teaching and
learning making work and research fun and pleasurable. He is remarkably modest,

never treated himself as superior and always took the time to understand and to
correct what I, a blabbering know-it-all kid with broken-english from half the way
across the world, is trying to say. Professor Mitsos is more than just an advisor, he
is my mentor in most aspects of life; even after four year, till this very moment, I am
still learning from him.

I would like to thank my thesis committee members, Professor Ghoniem and Pro-

fessor Buonjiorno, for their help, concerns, and smart feedback. Professor Ghoniem
has remarkable knowledge in the fields of science and engineering and even more
impressive is his ability to relate his knowledge and experience and to guide both

amateur and experienced researchers; all while being very humble and supportive.
Professor Buonjiorno is very critical and gives you all his attention and concerns
when you need him. He was able to anticipated pitfalls and point out important

issues.

Thanks to ENEL and the DOE and their research teams for sponsoring most of
this research and for their constant interest and feedback. Thanks for Aspen for

providing AspenPlus@ free of charge for academic uses.

I would like to thank my parents, for it was originally their dream that I pursue
a Ph.D. Also, because of my parents I was never worried or afraid about my future;

not because they were supportive or reasonable, but mainly because they were always

more worried and concerned and emotionally involved in my pursuit than I was. At

several occasions I was comforting them instead of the other way around; one can say
that they handled/lived the stressful emotional aspects of being a graduate student,

leaving the bare minimum for me. I thank my brilliant brother, who I greatly admire,

7
for being the voice of reason and for his constant constructive advice. I thank my
wonderful sister for being the most calming and relaxing person I have ever known,
and for always managing to alleviate any problem or concern.
Finally, thanks to my lab-mates and friends, they made my stay at MIT fun and

entertaining, and way different than the stereotypical ideas I had about the lifestyle
at MIT.

Disclaimer

This work was partly prepared as an account of work sponsored by an agency of the
United States Government. Neither the United States Government nor any agency
thereof, nor any of their employees, makes any warranty, express or implied, or as-
sumes any legal liability or responsibility for the accuracy, completeness, or usefulness

of any information, apparatus, product, or process disclosed, or represents that its use
would not infringe privately owned rights. Reference herein to any specific commercial

product, process, or service by trade name, trademark, manufacturer, or otherwise


does not necessarily constitute or imply its endorsement, recommendation, or favoring

by the United States Government or any agency thereof. The views and opinions of
authors expressed herein do not necessarily state or reflect those of the United States

Government or any agency thereof.


This thesis is in part based on [28, 29, 38, 35, 671.

8
This doctoral thesis has been examined by a Committee of the
Department of Mechanical Engineering as follows:

Professor Alexander Mitsos ..........................................


Thesis Supervisor
Visiting Scientist

Professor Ahm ed Ghoniem ...........................................


Member, Thesis Committee
Ronald C. Crane (1972) Professor

Professor Jacopo Buonjiorno.........................................


Member, Thesis Committee
Associate Professor of Nuclear Science and Engineering
10
Contents

1 A Double-Pinch Criterion for Regenerative Rankine Cycles 23

1.1 Summary . . . . . . . . . . . . . . . . . . 23
1.2 M otivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

1.3 Analytical Proof of Double-Pinch for Shortcut Methods . . . . . . . . 27

1.3.1 Two possibilities for pinch to occur . . . . . . . . . . . . . . . 27

1.3.2 Analytical Proof of Necessity . . . . . . . . . . . . . . . . . . 29


1.3.3 Graphical Proof of Uniqueness and Sufficiency . . . . . . . . . 36
1.3.4 Procedure for Cycle Optimization . . . . . . . . . . . . . . . . 38
1.4 Other Feedwater Configurations . . . . . . . . . . . . . . . . . . . . . 42

1.4.1 Drain to Open Feedwater Heater . . . . . . . . . . . . . . . . 42

1.4.2 Cascading (Downwards) . . . . . . . . . . . . . . . . . . . . . 42

1.4.3 Pumping to Feed . . . . . . . . . . . . . . . . . . . . . . . . . 43

1.4.4 Open Feedwater Heater . . . . . . . . . . . . . . . . . . . . . 46

1.5 Numerical Examples with a Simple Flowsheet . . . . . . . . . . . . . 46

1.5.1 Single Feedwater heater . . . . . . . . . . . . . . . . . . . . . 47

1.5.2 Non-Cascading, Cascading, and Common Practice . . . . . . . 56


1.6 Numerical Case Study with a Realistic Cycle Design . . . . . . . . . . 59
1.7 Conclusion and Future Work . . . . . . . . . . . . . . . . . . . . . . . 62

2 Optimal Design and Operation of Pressurized Oxy-Coal Combustion


with a Direct Contact Separation Column 65

2.1 Sum m ary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

2.2 M otivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

11
2.3 Flowsheet and Model Description .. .. .. .. . .. .. .. .. 69
2.3.1 Power Plant Flowsheet . . . . . . . . . . . . . . . . . 69
2.3.2 Flue Gas Pressure Losses . . . . . . . . . . . . . . . . 70
2.4 DCSC Modeling . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.4.1 DCSC Flowsheet . . . . . . . . . . . . . . . . . . . . 73
2.4.2 DCSC Operation . . . . . . . . . . . . . . . . . . . . 77
2.5 Optimization Formulation . . . . . . . . . . . . . . . . . . . 78
2.5.1 Objective Function . . . . . . . . . . . . . . . . . . . 78
2.5.2 Optimization Variables and Constraints. . . . . . . . 78
2.5.3 Integer Variables . . . . . . . . . . . . . . . . . . . . 81
2.5.4 Parameters Considered constant . . . . . . . . . . . . 82

2.5.5 Active Constraint Optimization . . . . . . . . . . . . 82


2.6 Results . . . . . . . . . . . . . . . . . . . . . . . 84

2.6.1 Variables at Optimal Operation . . . . . 86


2.6.2 Flue Gas Pressure Losses . . . . . . . . . 88
2.6.3 Capital Cost Reduction . . . . . . . . . 89
2.6.4 Validation of the Optimization Results . . . . . . . . . 90
2.7 Model-Based Optimization and Effect of Design Assumptions . 91

2.8 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

3 Pressurized Oxy-Coal Combustion: Ideally Flexible to Uncertain-


ties 99
3.1 Summary ................... 99
3.2 M otivation ................ 100

3.3 Flowsheet and Model Description . . . 101

3.4 Optimization Formulation . . . . . . . 102

3.4.1 Optimization Objective . . . . . 102

3.4.2 Design and Operation Variables 103


3.4.3 Constraints . . . . . . . . . . . 106
3.4.4 Active Constraint Optimization 106

12
3.5 Ideally Flexible Process to Coal, FWHs Areas, Input Flows and Spec-
ifications, and Ambient Temperature . . . . . . . . . . . . . . . . . . 110

3.5.1 Methodology for Flexibility Assessment . . . . . . . . . . . . . 110

3.5.2 Hierarchical Optimization . . . . . . . . . . . . . . . . . . . . 114


3.5.3 Flexibility to Input Flows and Parameters, (Air Flow, Slurry
water Flow, atomizer Stream Flow, and Oxidizer Stream Oxy-

gen Purity) . . . . . . . . . . . . . . . . . 123

3.5.4 Flexibility to Ambient Conditions . . . . . 126

3.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . 128


3.7 Future W ork . . . . . . . . . . . . . . . . . . . . . 131

4 Pressurized OCC Process Ideally Flexible to the Thermal Load 137

4.1 Summary ...................... . . . . . . . . . . . . . 137


4.2 Motivation .................. . . . . . . . . . . . . . 138
4.3 Turbine Performance Curves . . . . . . . . . . . . . . . . . . . . . 140

4.4 Modeling Approach . . . . . . . . . . . . . . . . . . . . . . . . . . 144

4.4.1 Process Operating Parameters . . . . . . . . . . . . . . . . 144

4.4.2 Flue Gas Pressure Losses . . . . . . . . . . . . . . . . . . . 146


4.5 Optimization Formulation . . . . . . . . . . . . . . . . . . . . . . 149
4.5.1 Design and Optimization Variables . . . . . . . . . . . . . 150
4.5.2 Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4.5.3 Active Constraint Optimization . . . . . . . . . . . . . . . 157
4.6 Results and Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 158
4.6.1 Flexibility Assessment . . . . . . . . . . . . . . . . . . . . 159
4.6.2 Behavior of Key Variables . . . . . . . . . . . . . . . . . . 164
4.6.3 Standard Rankine Cycles Without Pressurized Recovery 170
4.6.4 Partload and Subcritical Operation . . . . . . . . . . . . 173
4.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

5 A Split Concept for HRSG with Simultaneous Area Reduction and

Performance Improvement 177

13
5.1 Summary ..... ...................... 177
5.2 Motivation .............. ........ 178
5.3 Novel Split Concept . . . . . . . . . . . . . . . . . 180
5.3.1 Concept Description . . . . . . . . . . . . 180
5.3.2 Stand Alone HRSG-Split Simulation . . . 182
5.4 Optimization Formulation for Minimal Area and/or Minimal Compen-
sation Power Requirements . . . . . . . . . . . . . 188
5.4.1 Objective Functions . . . . . . . . . . . . 188
5.4.2 Optimization Variables . . . . . . . . . . . 194
5.4.3 Optimization Constraints . . . . . . . . . 194
5.4.4 Pareto Front Construction . . . . . . . . . 196
5.5 Results . . . . . . . . . . . . . . . . . . . . . . . . 196
5.6 Flexibility to Uncertainties . . . . . . . . . . . . . 199
5.7 Other Applications . . . . . . . . . . . . . . . . . 202
5.8 Conclusion . . . . . . . . . . . . . . . . . . . . . . 203

A Reaction Chemistry Added to the Separation Column in the DCSC


flowsheet 207

B DCSC Recirculation Water, rhRw-sr-in, Optimality Criterion 209

14
List of Figures

1-1 Pinch diagram for illustrative feedwater heater . . . . . . . . . . . . . 28

1-2 Pinch diagram demonstrating uniqueness of double pinch . . . . . . . 38

1-3 Feedwater configurations with pumping of the drain . . . . . . . . . . 43

1-4 Flowsheet for the numerical validation of the double pinch criterion . 48

1-5 Contours of efficiency for a fixed regenerated duty . . . . . . . . . . . 50

1-6 Contours of entropy generation rate Sge, for a fixed regenerated duty 52

1-7 Contours of efficiency for a fixed FWH Area . . . . . . . . . . . . . . 54

1-8 Contours of entropy generation Sen for a fixed FWH Area . . . . . . 55

1-9 Optimal performance of four different design procedures versus total


FW Hs area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

1-10 Realistic cycle design with four closed feedwater heaters . . . . . . . . 60

2-1 Oxycombustion cycle flowsheet based on wet recycling . . . . . . . . 72

2-2 Direct Contact Separation Column (DCSC) operation unit . . . . . . 75

2-3 Optimization variables and constraints for the pressurized OCC process 95
2-4 RHE and DCSC pressure parametric optimization and pressure para-

metric sensitivity result . . . . . . . . . . . . . . . . . . . . . . . . . . 96

2-5 Effect of design considerations . . . . . . . . . . . . . . . . . . . . . . 98

3-1 Evaluations performed for flexibility assessment . . . . . . . . . . . . 113

4-1 Variables and constraints for the pressurized OCC cycle for uncertainty
in load ........ ................................... 145

4-2 Results for flexibility to thermal load . . . . . . . . . . . . . . . . . . 163

15
5-1 HRSG single-split as part of a pressurized OCC process with a thermal
recovery unit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

5-2 Temperature profiles of four different operations of the flue gas . . . . 186

5-3 Results of HRSG-split multi-objective optimization . . . . . . . . . . 197

16
List of Tables

1.1 Specifications of flowsheet with 4+1 FWHs in Figure 1-10 . . . . . . 60

1.2 Results of flowsheet with 4+1 FWHs in Figure 1-10 . . . . . . . . . . 61


1.3 The applicability of the proposed design criterion for various configu-
rations and the evidence given in this chapter. . . . . . . . . . . . . . 63

2.1 Fixed Simulation Parameters for the pressurized OCC process . . . . 71

2.2 Optimization Variables for the pressurized OCC utilizing a DCSC . . 80

2.3 Optimization Constraints for the pressurized OCC process utilizing a

DCSC ........ ................................... 81


2.4 Optimization Results of the pressurized OCC utilizing a DCSC . . . . 97

3.1 Specifications of utilized coals . . . . . . . . . . . . . . . . . . . . . . 102

3.2 Design and operation variables facing fuel uncertainty . . . . . . . . . 107

3.3 Optimization constraints facing fuel uncertainty . . . . . . . . . . . . 108

3.4 The runs performed to check the flexibility of the OCC cycle, with an

RHE or DCSC thermal recovery unit, under fuel uncertainty . . . . . 112

3.5 Results for RHE flowsheet fuel flexibility A . . . . . . . . . . . . . . . 115

3.6 Results for RHE flowsheet fuel flexibility B . . . . . . . . . . . . . . . 116

3.7 Results for DCSC flowsheet fuel flexibility A . . . . . . . . . . . . . . 117

3.8 Results for DCSC flowsheet fuel flexibility B . . . . . . . . . . . . . . 133

3.9 Results for RHE flowsheet fuel and area flexibility . . . . . . . . . . . 135

3.10 Results for RHE flowsheet fuel, area, and ambient temperature flexibility136

4.1 Turbine performance data . . . . . . . . . . . . . . . . . . . . . . . . 143

17
4.2 Recycling pipes diameters and gas velocity ranges . . . . . . . . . . . 147
4.3 Design and operation variables for uncertainty in load . . . . . . . . . 155
4.4 Optimization constraints for uncertainty in load . . . . . . . . . . . . 156
4.5 Summary of results for RHE flowsheet fuel, area, and ambient temper-

ature flexibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160


4.6 60% part load flexibility . . . . . . . . . . . . . . . . . . . . . . . . . 161
4.7 30% part load flexibility . . . . . . . . . . . . . . . . . . . . . . . . . 162

5.1 Fixed input parameters for the HRSG-split . . . . . . . . . . . . . . . 187


5.2 Optimization variables for the HRSG-split . . . . . . . . . . . . . . . 195
5.3 Optimization results for the HRSG-split . . . . . . . . . . . . . . . . 200

A.1 The relevant reaction added to Separation Column of the DCSC flowsheet208

18
Nomenclature

Latin symbols
PB Bleed pressure [bar] izcp Capacity rate [kW/K]
Derivative of the extrac- Derivative of the extrac-
"PB
0 tion mass flowrate follow- ) tion mass flowrate fol-
ing pinch at the bleed out- lowing pinch at the on-
let with respect to extrac- set of condensation with-
tion pressure [kg/(s bar)] respect to extraction pres-
sure [kg/(s bar)]
hT(Po) Enthalpy at the turbine hB,o Enthalpy of the bleed
outlet [kJ/kg] stream in the outlet of the
feedwater heater [kJ/kg]
hgsat Enthalpy of the saturated A Heat Transfer Area [m 2 ]
vapor [kJ/kg]
U Heat Transfer Coefficient rh Mass flowrate [kg/s]
2
[kW/(m K)]
PO Outlet pressure of the tur- P Pressure [bar]
bine
Sgen Rate of entropy generation Q Regenerated Duty
[kW/K] [kW, MW]
CIP Specific liquid thermal ca- T Temperature [C, K
pacity [kJ/(kg K)]
WB Power that the extraction
stream would have gener-
ated in the turbine [kW]
Pcomb Combustion Pressure (bar) PDeaerator Deaerator pressure (bar)
Q Duty Transfer (MW) T Temperature (OC)
Qcomb Combustor Duty (MW) qDeaerator Quality in Deaerator tank
Greek symbols
r_ Cycle Efficiency (%)
Acronyms
CFWH Closed Feed Water Heater FWH Feed Water Heater
HX Heat Exchanger MITA Minimum Internal Temper-
ature Approach [C, K]

19
OFWH Open Feed Water Heater o-pinch Pinch at bleed exit from
FWH
p-pinch Pinch at onset of bleed con- TLoad Thermal Load
densation
ASU Air Separation Unit CCS Carbon Capture and Se-
questration
FG-Rec-pri Primary Recycled Flue Gas FG-Rec-sec Secondary Recycled Flue
I _Gas
Cool-Gas Flue Gas at exit of HRSG CSU Carbon Sequestration Unit
DCSC Direct Contact Separation DCSC-HX Heat Exchanger in DCSC
Column thermal recovery
unit
CPR Compensation power re- CER Compression enthalpy rise
quirements to overcome due to the CPR (kW)
flue gas pressure losses
(kW)
HRSG Heat Recovery Steam Gen- RHE Recovery Heat Exchanger.
erator Acid condensation occurs
in RHE
FG-DCSC-in Flue Gas entering the FG-RHE-in Flue Gas entering the RHE
DCSC
FW Feedwater FW,Main Main feedwater stream.
Largest working fluid flow
passing through the HRSG
and entering the expansion
line
FWH Feedwater Heater FW-HRSG-in Feedwater entering the
HRSG. Same flowrate as
FW, main
FW-DCSC-in Feedwater entering the HHV Higher Heating Value of
DCSC Coal (MJ/kg)
Hot-Gas Flue gas entering the HP-Pump High Pressure Pump
HRSG
LHV Lower Heating Value of BLD Rankine cycle regeneration
Coal (MJ/kg) bleed
LP-Pump Low Pressure Pump OC Oxy Combustion
OCC Oxy-Coal Combustion RW-Sep-in Recirculation water of the
DCSC entering the separa-
tion column
Cool-Gas Flue gas exiting the HRSG FW-HRSG-in Feedwater entering the
HRSG
Comb-Gas-in Flue gas entering the com- FW-Recov-out Feedwater exiting the re-
bustor covery unit, RHE/DCSC
BLD1-stage Bleedi extraction stage BLD2-stage Bleed2 extraction stage
BLD3_stage Bleed3 extraction stage C02_Cap Ratio of CO 2 capture to to-
tal produced

20
C02...pure Purity of CO 2 captured MITA-FWHi Minimum internal temper-
ature approach of FWHj
(OC)
MITA.HRSG HRSG pinch ('C) MITA-RHE RHE pinch ('C)
PBLDi Bleeds extraction pressure PComb Combustion Pressure (bar)
(bar)
PDeaerator Deaerator pressure (bar) QFWHi Feedwater Heater i duty
I (kW)
OComb Combustor Duty (MW) qDeaerator Quality in Deaerator tank
TCool-Gas Temperature of flue gas ex- TFG-RHE-out Flue gas temperature at
iting the HRSG ('C) RHE exit (OC)
TFW-HRSG-in Temperature of feedwater TComb-Gas-in Temperature of gas enter-
entering the HRSG ('C) ing the combustor(OC)
Subscripts
a actual design and operating b basecase design and operat-
conditions ing conditions
B Bleed B, i Bleed in
B, o Bleed out F Feed
F,i Feed in F,o Feed out
o Following o-pinch p Following p-pinch
T Turbine expansion line
Superscripts
g Vapor 1 Liquid
lg Evaporation sat Saturated state
Pressure Drop Parameters
Ac HRSG cross section area As HRSG total surface area
(m2) (m2)
d Recycling pipe diameter D HRSG tube diameter (in)
(in)
Dh Hydraulic diameter (in) ATim Log mean temperature dif-
ference (K)
APHRSG HRSG pressure drop (Pa) APipe Recycling pipe pressure
drop (Pa)
6 Wall roughness (m) f Friction factor
F ATm correction factor H HRSG height (i)
k Thermal conductivity of L HRSG length (m)
flue gas (W/m.K )
Lpipe Recycling pipes equivalent r Flue gas flowrate (kg/s)
length (m)
A Dynamic viscosity (kg/m.s) N Number of tube rows along
HRSG length
Nu Nusselt Number QHRSG Total transferred duty in
I_ HRSG (W)
Re Reynolds number p Density (kg/m 3 )

21
SL HRSG longitudinal pitch ST HRSG transverse pitch (m)
(m)
U HRSG overall heat transfer V Bulk flue gas velocity (m/s)
coefficient (W/m 2 K) I
Vo Average gas velocity at Vmax Maximum gas velocity in
HRSG entrance (m/s) HRSG (m/s)
W HRSG width (m) I I

22
Chapter 1

A Double-Pinch Criterion for

Regenerative Rankine Cycles

1.1 Summary

A double-pinch criterion for closed feedwater heaters (FWH) of regenerative Rankine

cycles is presented. The FWHs are modeled as counter-current heat exchangers.

Thus, two potential pinch positions in the FWH exist: (i) at the exit of the bleed

(drain) and (ii) at the onset of condensation. For a given heat duty in the FWH, feed

inlet temperature and flowrate, the extraction flowrate and pressure should be chosen

to achieve the same minimal approach temperature at the two potential pinch points.

An analytical proof is given for a fixed pinch value for the case that the drain enters

the condenser, based on weak assumptions. Additionally, the criterion is numerically

demonstrated for fixed pinch value and for fixed heat exchanger area using the most

common configurations: drain to condenser, drain to deaerator, and drain cascaded

to next FWH. A similar criterion is developed for the case that the drain is pumped

(upwards or downwards) and mixed with the feedwater. The double pinch criterion

simplifies the optimization procedure and results in significant efficiency increase for

23
fixed heat exchanger area. For numerical reasons it is advisable to use the pressure

as the optimization variable and calculate the heat duty and mass flowrate.

1.2 Motivation

Rankine cycles are widely used in power generation, typically with features for effi-

ciency increase such as reheat, super heat, and regeneration [1, 2, 3]. Commercial

software packages capable of high fidelity modeling of the power cycles are available,

e.g., Thermoflex@ [4], AspenPlus@, GateCycle@ [5]. These software packages offer

tools for performing parametric and optimization studies on given cycles (flowsheet

connectivity) or even construct power cycles for a given application.

All but the simplest cycles are regenerative, i.e., they include preheating of the feed

(return from the condenser) via extraction of bleed streams from the turbine [1, 2, 3].
This preheating is performed in closed and/or open feedwater heaters (CFWH and

OFWH). There are various configurations of FWHs depending on where the exit bleed

of the CFWH, i.e., the drain, is sent to. The performance of the power cycle increases

in general with the number of FWHs. Thus, typically the number of FWHs is selected

based on cost considerations, i.e., balancing capital and operating costs. A well know

approximate design criterion is to have an equal enthalpy rise across each FWH in the

regeneration section of the Rankine cycle. More precisely, under some idealizations,

for maximum efficiency in a non-reheat non-supercritical plant the enthalpy rise of

the feedwater up to the point of saturation should, to a first approximation, be the

same in all heaters and the economizer [1, 2].

CFWHs are essentially multi-phase heat exchangers (HX). The design of HX net-

works is a well-established field, e.g., [6, 7, 8, 9, 10, 11]. The design of both power

cycles and HX networks can be performed with either shortcut methods or more rig-

orous models. A very common shortcut method is the so-called pinch analysis, i.e.,

24
to select a minimal temperature approach (MITA) and then calculate the inlet condi-

tions and heat duty for each HX. This shortcut has the advantage of decoupling the

capital costs and detailed design of the HXs from the operating costs. The implicit

assumption of the pinch method is that the required HX area is accurately charac-

terized by the MITA. However, accurate calculation of heat transfer area is needed

for economic optimization.

The focus of this chapter is to propose a new criterion for the optimization of

regenerative power cycles and apply it to both pinch method and optimization for a

fixed heat transfer area. Throughout the chapter CFWHs will be treated as counter-

current HXs. Consequently, there are two possible pinch points, namely at the onset of

condensation and at the exit of the bleed (the drain). The main result herein is that for

most FWH configurations optimal operation is equivalent to simultaneously achieving

both pinches. In all considerations the connectivity between units (flowsheet) and

the expansion line (condenser temperature, turbine inlet pressure & temperature and

total flowrate) are kept fixed. For the working fluid a pure species with phase change

is considered.

In the pinch analysis, a MITA is selected using economic criteria as a surrogate for

HX area. Subsequently, there are three degrees of freedom for each CFWH, namely

heat duty (as a surrogate via the MITA), bleed extraction pressure (and thus tempera-

ture) and bleed extraction flowrate. Thus, at least in principle, the cycle performance

can be optimized numerically considering simultaneous variation of these degrees of

freedom, subject to the constraints of minimal approach temperature. In the rigorous

analysis, heat transfer area is selected via economic analysis; the remaining degrees

of freedom are bleed extraction pressure and bleed extraction flowrate; moreover, the

MITA is free. The proposed design criterion eliminates two of the variables for each

CFWH. Moreover, for the case of the pinch analysis, the proposal eliminates the need

to check for the pinch-violation constraints and as such the need for a spatially dis-

25
tributed model for the feedwater heater. In summary, the proposal simplifies the cycle

optimization drastically. This computational acceleration is particularly important

when Rankine cycles are not considered in isolation, but rather as a part of a com-

plicated process, e.g., oxycombustion [241, or inside another procedure, e.g., selection

of working fluid in an organic Rankine cycle. In fact the identification of the crite-

rion was achieved in the process of optimizing a pressurized oxycombustion cycle [24]

using deterministic local and heuristic global optimization methods. Optimization of

a Rankine cycle in isolation is relatively simple, but not trivial in general-purpose

modeling tools, which in the absence of the proposed criterion result in suboptimal

local optima. The criterion also enables a simpler use of approximate design criteria,

such as the aforementioned equal enthalpy rise across feedwater heaters.

In Section 1.3 the shortcut rule of minimum temperature approach is considered

for the simplest FWH arrangement, namely that the drain is sent to the condenser.

A precise statement is given with appropriate boundary conditions and proved ana-

lytically. Moreover, it is demonstrated how the criterion can be implemented inside

an optimization procedure resulting in a significantly simpler optimization formula-

tion than the alternative of simultaneous optimization of all variables. In Section 1.4

the applicability of the criterion to other configurations is discussed. In Sections 1.5

and 1.6 case studies with various FWH configurations are calculated numerically both

for the shortcut calculation and for a constant area respectively, discussing also the

entropy generation in the feedwater. The proposed criterion results in significant sav-

ings compared to current design practice for both shortcut and rigorous calculation.

26
1.3 Analytical Proof of Double-Pinch for Shortcut

Methods

In the following, the double pinch design criterion is developed and proved based on

a set of nonrestrictive assumptions. First, it is demonstrated that only two points in

the pinch diagram are of interest, namely the onset of condensation and the outlet

of the bleed. Subsequently, for the case that the drain is sent to the condenser,

it is proved that for a given heat duty a double pinch is necessary for optimality.

Moreover, there exists a unique pair of extraction pressure and flowrate that results

in a double pinch. Therefore, a double pinch is also sufficient for optimality. Then, a

reordering of variables is proposed along with a procedure for computationally efficient

optimization.

1.3.1 Two possibilities for pinch to occur

Assumption 1.3.1 (Capacity Rates). The ratio of feed flowrate to extraction flowrate

is assumed sufficiently high that the capacity rate (rhcp) of the feed is higher than that

of the bleed for both the superheated and subcooled regions.

Assumption 1.3.1 holds for typical Rankine cycles.

The following proposition shows that in the pinch diagram only two points are of

interest. For a graphical illustration, compare also Figure 1-1.

Proposition 1.3.2. Under Assumption 1.3.1 a minimum approach temperature be-

tween the feed and bleed can only occur at two points, namely the onset of condensa-

tion, and the bleed outlet.

Proof. Recall that the feedwater heater is modeled as a counter-current HX. The feed

is always subcooled and therefore a smooth curve is obtained in the pinch diagram.

In contrast, the bleed consists in general of three regions, namely, the superheated

27
region, the condensation region, and the subcooled region, giving two kinks at the

transitions. The region of condensation results in a horizontal line for the bleed,

i.e., enthalpy decrease without temperature change. By Assumption 1.3.1, the other

two curves corresponding to the bleed have a higher slope than the slope of the curve

corresponding to the feed. Therefore, in the direction of flow of the bleed from the inlet

to the outlet, the superheated and subcooled regions result in convergence between

the bleed and the feed curves, whereas in the condensation divergence between the

two curves is observed. Consequently, the two potential points for a pinch are the

onset of condensation and the bleed outlet. Note that if the bleed inlet is in the

two-phase region, the onset of condensation coincides with the inlet. E

300 ________ I_______I__


Bleed M=25.5kg/s P=25bar
Bleed M=30.Okg/s P=1 5bar
250 Feedwater M=100.Okg/s P=100bar

200

c 150 p-pinch
E
I-
100 --

50 -

0-pinch-
01'
0 1 2 3 4 5 6 7 8
Thermal Energy Transfer, Duty (kJ/s) X 104

Figure 1-1: Pinch diagram for illustrative feedwater heater (results generated in
AspenPlus@). There are only two possibilities for pinch to occur in a heat regenera-
tion in the FWH of a pure substance; at the cold end of the heat exchanger (drain, red
dotted bleed line, labeled o-pinch), or at onset of bleed condensation (black dashed
bleed line, labeled p-pinch).

28
1.3.2 Analytical Proof of Necessity

In this section the shortcut method of minimum temperature approach is considered

for the case that the drain is sent to the condenser. It is shown that for a given feed

flowrate rnf, feed inlet temperature T,j and heat transfer duty Q a double pinch is
optimal except for trivial cases. It is based on relatively weak assumptions.

Assumption 1.3.3 (Bleed saturation enthalpy). It is assumed that the enthalpy of

the saturated vapor hg,' at the bleed pressure PB is not lower than the enthalpy at

the turbine outlet hT(Po).

h9a'(PB) > hT(P,),

where the subscript T denotes the turbine (expansion line) and Po the outlet pressure

of the turbine.

Assumption 1.3.3 is satisfied for typical expansion lines. It could only be violated if

the turbine outlet state is highly superheated and the heat duty is very small allowing

for a very low extraction pressure PB. Such an operation is suboptimal.

Assumption 1.3.4 (Bleed outlet enthalpy). It is assumed that the enthalpy of the

bleed stream in the outlet of the feedwater heater hB,0 is not higher than the enthalpy

of the turbine outlet hT(Po)

hB,o <; hT(Po).

Assumption 1.3.4 is satisfied for typical expansion lines and working fluids, e.g.,

water and ammonia, heptane and toluene.

Assumption 1.3.5 (Bleed pressure). It is assumed that the optimal extraction pres-

sure PB is strictly higher than the turbine outlet pressure P.

Assumption 1.3.5 can only be violated if the heat duty of the feedwater heater is

extremely low and the outlet of the turbine is highly superheated. In other words,

recuperators are not considered herein.

29
Assumption 1.3.6 (Bleed pressure). It is assumed that the optimal extraction pres-

sure PB is strictly lower than the turbine inlet pressure.

Assumption 1.3.6 holds for typical Rankine cycles.

Assumption 1.3.7 (Heat capacity and saturated vapor). The following relationship

is assumed to hold for the pressures of interest

&hg,sat &Tsat
<; c<P(T"(PB) - AMITAT, PF) -
aP PB OP PB

Assumption 1.3.7 is satisfied for typical working fluids, such as water, toluene and

ammonia. For the conditions of interest, the expression of the left hand side is either

negative or slightly positive, whereas the right hand side is always positive.

Assumption 1.3.8 (Positive Pressure Dependence of Liquid Enthalpy). The deriva-

tive of the liquid enthalpy with respect to pressure is positive

-h > 0.
aki T

Note that Assumption 1.3.8 holds for an incompressible liquid, which is a good

approximation for the liquids. Moreover, it holds for typical working fluids, such as

water, toluene and ammonia.

In the formal statement of the main result, reheats are excluded for simplicity.

Moreover, a fixed expansion line is considered, essentially assuming that the turbine

isentropic efficiency is not affected by the extraction conditions.

Theorem 1.3.9 (Double pinch necessary). Consider a regenerative Rankine cycle

without reheats and with positive isentropic efficiency of the turbine. Let the turbine

inlet temperature, inlet pressure, outlet temperature and outlet pressure P be fixed,

i.e., not influenced by the extraction. Consider an arbitraryfeedwater heater, specified

by an arbitrary but fixed feed flowrate rilF, feed inlet temperature TFj, and a heat

30
transfer duty Q. Suppose that the feed pressure Pf is chosen so that the feed stream

remains subcooled. Suppose that the feedwater heater can be modeled as a counter-

current heat exchanger with a minimum approach temperature AMITAT and without

pressure drop. Assume that the extraction state is saturated or superheated, i.e., the

extraction temperature is not smaller than the saturationtemperature of the extraction

pressure. Suppose that the drain is sent to the condenser. Select a pair of extraction

flowrate rmB and pressure PB such that the cycle performance is optimized. Under

Assumptions 1.3.1 through 1.3.8 a double-pinch occurs, i.e., the MITA occurs both at

the condensation onset and at the outlet of the bleed stream.

Proof. The proof is done by contraposition, i.e., by considering an optimal pair nB,

PB, assuming that no double pinch occurs and concluding that the pair is not opti-

mal. This is done in three steps, first excluding the case that no pinch occurs, then

excluding the case that a pinch occurs only at the onset of condensation and finally

excluding the case that a pinch occurs only at the bleed outlet.

Based on the assumption that no reheat exists, the power that the extraction

stream would have generated in the turbine is given by

WB = r4B(hT(PB) - hT(Po)). (1.1)

Note that hT(PB) is the enthalpy of the bleed inlet to the feedwater heater hT(PB) =

hB,i. Optimal extraction minimizes this lost power WB.

Note first that if the extracted steam is saturated steam or in the two-phase

region (hT(PB) h9,,at(pB)), then the pinch at the onset of condensation is trivially

satisfied. Moreover, the pinch at the outlet minimizes ThB thus minimizing WB for a

given pressure since PB > P0 is assumed. Thus, we can assume hT(PB) > h9,St(PB)-

1. Suppose for contraposition that the MITA is not reached in the feedwater for

(PB, MB). Then, an infinitesimal reduction of either extraction flowrate or pres-

31
sure (or even both simultaneously) allows the same heat transfer duty without

a violation of the MITA. On the other hand, this reduction implies that more

power is produced in the turbine, see (1.1), or the original pair is not optimal.

2. Assume now that for (PB, MB) we have a pinch at the onset of condensation

but not at the bleed outlet. Maintaining the pinch at the onset of condensation

implies that a change in extraction pressure results in a change in extraction

flowrate. This is possible without violating any constraints, given Assump-

tion 1.3.5. We denote the derivative of the extraction mass flowrate following

pinch at the onset of condensation with-respect to extraction pressure

where the subscript p stands for p-pinch.

Consider the partial derivative of the power not produced in the turbine with re-

spect to the extraction pressure following the pinch at the onset of condensation

evaluated at (PB, TnB)

B 0hB - Oh\ (1.2)


=(hT(PB) - hT(Po)) + 7B (9.PB
9PB - =\OPB -B ± PB
PB PBn

where we have used the fact that the turbine inlet and outlet states are fixed.

Optimal operation implies minimal WB, or if we find that a- 0 then the


PB
pair (PB, MB) is not optimal.

In the right-hand side of Equation (1.2) the second term is positive (f > 0),
since the turbine produces work. Thus, if the derivative evaluated at

PB is nonnegative, we also have B > 0 or the pair (PB, MB) is not optimal.

Note that this would imply that we can reduce both extraction pressure and

flowrate and maintain the same preheating. Note also that (B) < 0 can be

directly proved.

32
We thus only need to consider the case

OrhB <0.
OPB / (1.3)
<-

We will show that the derivative of the power lost with respect to the
extraction
pressure is negative, or increasing the extraction pressure increases
efficiency.

We will generate an expression for (0-) in the following. The condition for
a pinch at the onset of condensation is

rnF (hFO - h(T sat(PB) - AMITAT, PF rB (hT(PB) - h9 at(PB))-

Taking the derivative with respect to PB and evaluating at (PB, MB)


gives

-~FCP(Tsat(PB) - AMITAT, PF)


OP IPB
C
rnB
(hT(PB) - h9,sat (pB)) + MB OhT - hgB
t
PB MB O P PB

or

(hT(PB) - h t (PB)
Bsa
\&PB )

- ih T +- Ohgsat "Tsat
= MB h +MB - hFCP(sat(PB) - AMITAT,PF) OP/(1.4)
5 PB PB
PB

By Assumption 1.3.3 we have hT(Po) < hg'8at(PB) and therefore

hT(PB) - hg,sa t (PB) h(P) --- hT(Po).

33
Recalling Inequality (1.3) we thus obtain

(hT(PB) - hT(Po)) (<hB ) (h(PB) - h'sat(PB)) OPB


PBP P

and therefore combining Equations (1.2) and (1.4) we obtain

WB- ,sat s (Ts*t


< +mB _- B FC-p(T ( PB) ~ AMITAT PF) P
PB PB PB

Noting that nB < rF by Assumption 1.3.7 we obtain '9WB


0 PB ,PB < 0. By As-
sumption 1.3.6 it is possible to increase the pressure and thus we have shown

that (PB, MB) is not optimal.

3. Assume finally that for (PB, MB) we have a pinch at the bleed outlet but not

at the onset of condensation.

Similarly to the previous case we will consider variation of the extraction pres-

sure by maintaining the pinch at the bleed outlet ( ), i.e., the derivative

of the extraction mass flowrate following pinch at the bleed outlet with respect

to extraction pressure, where the subscript o stands for o-pinch.

Equivalently to the previous case we obtain

09WB - (hB (hT(PB) - hT(Po)) + MB (1.5)


PB B PB B

and
(0mB) < 0. (1.6)
PB PB

We then show that the derivative of the power lost with respect to the extraction

is positive, implying that the extraction pressure should be decreased.

34
We will generate an expression for (2) in the following. The heat transfer

duty maintaining a pinch at the outlet can be calculated as

Q = rB (hT(PB) - h1 (Ti + AMITAT,PB))

Noting that the total heat transfer is constant, the derivative with respect to

PB is zero. Evaluating the derivative of the right hand side at (PB, MB) thus
gives

(hT(PB) - h'(T,i + AMITAT, PB)) ±MB -


0 = OPB O-B 9PB PB
SPB ;,i+AMITAT,PB

or

O(nB OhT Ohl


OPB / -
(hT(PB) - h'(Ti + AMITAT, PB)) B
PB ~
+mB OPB TT+AMITATPB

(1.7)

By Assumption 1.3.4 we have hl(Ti + AMITAT, PB) = hB,0 <! hT(P) and

therefore

hT(PB) - h'(Tri + AMITAT,PB) hT(PB) - hT(P).

Recalling Inequality (1.6) we thus obtain

(hT(PB) - hT(Po)) (0riB) h(PB) - h'(Ti + AMITAT,PB))


B PB
(
OB )
PB

and therefore by (1.7)

ohi
(hT(PB) - hT(Po)) B > -BMB +rnB
OPP
O PB T,I±AMTAT,PB

35
which together with (1.5) gives

9WB - B
B PB TT,i+AMITAT,PB

and by Assumption 1.3.8 we obtain > 0. By Assumption 1.3.5 it is

possible to decrease the pressure and thus we have shown that (PB, MB) is not

optimal.

1.3.3 Graphical Proof of Uniqueness and Sufficiency

Theorem 1.3.9 proves that a double pinch is a necessary condition for optimality. In

principle, it allows for multiple double pinches, out of which some may be suboptimal.

Under two additional assumptions it is possible to prove that for a given heat duty

there exists a unique pair (PB, rnB) that gives a double pinch.

Assumption 1.3.10 (Weak Pressure Dependence of Subcooled State). The deriva-

tive of the liquid enthalpy with respect to pressure is smaller than the derivative of the

enthalpy in the expansion line

O < hT
(1.8)
T,PB PB

Moreover, the heat capacity in the subcooled region is assumed to be a weak function

of pressure for the temperatures and pressures of interest, or more precisely for any

two pressures PB1, PB2 such that PBl> PB2 we have

c (T, PB1) hT(PB1) - h1(T, PBj)


<T(B)-hl(,P2 (1.9)
cip (T,PB2) h B2--hTB2

36
Both clauses of Assumption 1.3.10 hold unless the turbine efficiency is extremely

low.

Assumption 1.3.11 (Enthalpy of Vaporization Decreasing). The derivative of the

enthalpy of vaporization with respect to pressure is negative

&hg
< 0.

Assumption 1.3.11 holds for pure substances [12].

Lemma 1.3.12 (Double pinch unique). Consider the conditions and assumptions of

Theorem 1.3.9. Under the additional Assumptions 1.3.10 and 1.3.11 there exists a

unique pair of extraction flowrate 71B and pressure PB that gives a double pinch.

Proof. Consider a pair (PBi, rnBl) that results in a double pinch. Consider a second

pair PB2, rnB2 with PB2 < PBi that results in a pinch at the outlet. We will show

that it violates the minimal approach temperature at the onset of condensation.

Consider the pinch diagram, Figure 1-2, noting that no linearity is assumed (which

would require constant heat capacity in some regions). By Assumption 1.3.1 the slopes

of the subcooled curves are bigger than that of the feed for both bleeds.

Let the points Bli, B gj', B'sat and Bio denote respectively the inlet of the bleed,

the onset of condensation, the onset of subcooling and the outlet of the bleed. More-

over, let Bis"" denote the intersection of the saturation temperature at pressure PB2

with the subcooled curve of bleed 1. Finally, let Bg~mt denote the intersection of

the saturation temperature at pressure PB2 with a parallel to the subcooled curve of

bleed 1 going through Bfi"a.

For bleed 2 to achieve pinch at the outlet, its outlet coincides with Bi0 . Note that

37
pinch at the outlet implies

= nB (hT(PB1) - h1 (Ti + AMITAT, PB1))

Q =rnB2 (hT(PB2) - h'(T, i + A MITAT, PB2))

We will now employ the two inequalities in Assumption 1.3.10. By (1.8) the two

conditions for Q directly imply that rnB2 > rnBl. Moreover, by (1.9) the same

assumption, the onset of subcooling of bleed 2 is to the left of Beal. Finally, together

with Assumption 1.3.11, the onset of condensation is to the left of B l Recalling

that the slope of the bleed is higher than that of the feed, this point is to the left of

the minimum temperature approach.

Figure 1-2: Pinch diagram demonstrating uniqueness of double pinch

Since the bleed pressures PB1, PB2 are arbitrary, we can also exclude the case of

a double pinch with higher pressure than PB1 say PB3. Indeed, suppose that we have

a double pinch for the bleed pressure PB3. By the above arguments PB1 violates the

minimal approach temperature leading to a contradiction. E

Theorem 1.3.13. Under the assumptions of Theorem 1.3.9 and Lemma 1.3.12 the

unique pair (PB, rnB) that gives a double pinch is optimal.

The proof of Theorem 1.3.13 is trivial and is omitted. Note that we take existence

for granted; this is justified by the change of variables in the next subsection.

1.3.4 Procedure for Cycle Optimization

Theorems 1.3.9 and 1.3.13 prove that a double pinch is optimal for a fixed heat duty.

However, it is computationally more efficient to vary the extraction pressure in both

shortcut of pinch analysis and fixed area approach.

38
Procedure for Pinch Analysis

For the pinch analysis approach, it is possible to directly calculate the pair of extrac-

tion flowrate rnB and heat duty Q that leads to double pinch

. hi(Tsat(PB)- AMITAT, PF) - h'(TFi, PF)


mF hgsat(PB) - hi (TFi
+ AMITAT, PB)

Q = hB(hT(PB) - h(TFi ± AMITAT, PB)) (1.10)

Note that since the expressions are explicit in raB and Q there is a unique double

pinch for a given extraction pressure PB. One of the advantages of this change

of independent variable is that the explicit equations for the occurrence of pinches

eliminate the need for a spatially distributed model of the feedwater heater. However,

we have not proved whether a double pinch is optimal for a fixed pressure. It is

therefore necessary to demonstrate that the change in variables does not result in

convergence to spurious solutions when used inside an optimization algorithm. In the

following it will be shown that a global/local optimum in the pressure space following

the double-pinch implies a global/local optimization of the design and operation.

Proposition 1.3.14 (Extraction pressure as independent variable does not introduce

complications). Let the superscript k = 1,. . . ,n denote the feedwater k. Suppose that

optimization is performed with respect to the extraction flowrates Pk with the heat

duty Qk and extractionflowrate rhk specified by (1.10) and that Pk is found optimal

in a set 'pk. Denote the corresponding extraction flowrates rik and heat duties Qk.

If 'pk contains all possible extraction pressures, then the triplets (PB ,I ) ge

a globally optimal power cycle efficiency. If 'k is a neighborhood with Pk in the

interior,the triplets give a locally optimal power cycle efficiency.

Proof. Consider first the case that the sets 'Pk encompass all possible extraction pres-

sures. This implies that the triplets P, Mi, Qk) are optimal over all triplets leading

to a double pinch, and therefore by Theorem 1.3.9 also optimal over all triplets.

39
Consider now the case that 'Pk are neighborhoods with PL in the interior. Assume

first that there is a single feedwater heater k = n = 1. Let the solution to (1.10)

as a function of the extraction pressure Pk be denoted as rhkd(PA), Q(Pk). Local

optimality implies

S(P, rnBd(B, (p k k) v 'Pk,

The continuity of the mappings in (1.10) implies that the image of Qk on ',k is an
interval. Proposition 1.3.12 ensures a unique double pinch for a fixed heat duty and

thus Qk is not at the boundary of the interval. By Theorem 1.3.9 for any Pwe have

Thus, there exists a neighborhood with (Pk, ri, Q ) in the interior for which (Pt, i, Qk)

is optimal. This is the definition of a local minimum.

Assume now multiple FWHs for the case of local optimality. Recall that the

flowsheet has one or multiple points with fixed temperature, e.g., the condenser.

Move in direction of the feed and between each of these points divide the feedwater

heaters in pairs and if needed an additional feedwater heater. For the pairs consider

a variation of the extraction pressure Pk in a neighborhood containing PB interior

for the first and adjust the second P+1 such that the sum of the heat duties remains

constant 0k+1 + k = + Qk. Similarly to the case of a single feedwater heater


±k+I

we can use uniqueness and continuity (with respect to extraction pressure and feed

inlet temperature) to construct a neighborhood with Qk, Qk+1 in the interior and

prove local optimality. Given that the pair has a constant sum of heat duties, we can

treat the odd (separate) feedwater heater similarly to the case of a single feedwater

heater. 0

40
Procedure for Fixed Area

The double-pinch criterion can also be applied in the case of fixed heat transfer area,

but some iterative procedure is required. The optimal procedure somewhat depends

on the flowsheeting software used. There are several plausible choices, and here

only the more promising are discussed. The general recommendation is to have the

extraction pressure PB as the main optimization variable. Then, for a given heat

transfer area and extraction pressure the double-pinch criterion fully specifies the

operation of the FWH, by eliminating one variable, e.g., the bleed flowrate. Note

that the value of the pinch is an unknown and that the operation is only implicitly

specified.

The first main choice is to either let the optimizer control an additional variable

to satisfy a nonlinear constraint, or mask this pair from the optimizer and embed it as

a design specification inside the objective function evaluation. The former procedure

is recommended by the recent excellent treatment of chemical plants by Biegler [131.

However, in the computational experience herein and in [241, the use of embedded

design specifications was found more favorable, because it avoid failures at the simu-

lation level.

The second consideration is which pair of variable and constraint to select. Herein,

the bleed flowrate fTB is adjusted to meet the double-pinch criterion. This requires a

calculation of the heat transfer in the FWH for each iteration ("HX analysis"). Once

this calculation is performed, the values of the two pinch points can be obtained from

the temperature profiles or from solving (1.10). An alternative is to vary the value

of the pinch AMITAT to meet the given heat transfer area. In this alternative the

bleed flowrate and heat transfer duty can be calculated explicitly from (1.10) but the

calculation of the heat transfer area is required ("HX design").

41
1.4 Other Feedwater Configurations

As aforementioned there are several FWH configurations. For most configurations

the double pinch criterion seems plausible but it is outside the scope of this paper

to prove analytically. For some configurations the criterion is not applicable. For a

summary, see Table 1.3.

1.4.1 Drain to Open Feedwater Heater

In the above analytical proofs it was assumed that the drain is sent to the condenser.

For high-pressure CFWHs an obviously better choice is to send the drain to the next

possible deaerator or OFWH, since this allows the recovery of some of the remaining

availability and reduces the load on the condenser and the pumps. The proof employed

in Theorem 1.3.9 assumes that the drain does not affect the temperature of the feed

inlet and therefore is not directly applicable to the case that the drain is sent to the

condenser. A claim of this chapter is that the double pinch criterion is applicable to

this feedwater configuration. An analytical proof is outside the scope of this chapter

and instead numerical examples are given in Section 1.6.

1.4.2 Cascading (Downwards)

In high-efficiency cycles with multiple CFWHs it is customary to cascade the bleeds

downwards, namely send the drain to the next CFWH (immediately lower pressure)

and mix it with the bleed inlet. Similarly to sending to an OFWH some of the remain-

ing availability is captured. As demonstrated in numerical examples, Section 1.5 and

Section 1.6, the double pinch criterion is promising. However, the analytical proofs

given in Section 1.3 are not directly applicable, and an analytical proof is outside the

scope of this study.

42
1.4.3 Pumping to Feed

One flowsheet configuration is to pump the drain to the feed pressure and mix the

feed, see Figure 1-3. One alternative is to mix at the inlet of the CFWH, which is

referred to as pumping backwards (or downwards). The other alternative is to pump

forward (or upward), i.e., to mix with the feed at the outlet of the CFWH. For either

of the pumping configurations, double pinch is an optimal selection of bleed pressure

and flowrate for the pinch analysis but not the unique optimum. For the constant

area approach, double pinch is not advisable. In general, an optimal selection is to

achieve pinch at the onset of condensation and just enough subcooling at the outlet to

ensure that no technical difficulties arise for pumping. Similarly to the double pinch

criterion, this gives two constraints which can be used to eliminate two of the three

variables.

JFWH FWH

ump Pump

Figure 1-3: Feedwater configurations with pumping of the drain; left pumping back-
wards/downwards, right pumping forward/upwards.

Assumption 1.4.1 (Feed inlet enthalpy). It is assumed that the enthalpy of the feed

inlet to the feedwater heater hF'i is not higher than the enthalpy of the turbine outlet

hT(Po)

hFi <; hT(Po).

Assumption 1.3.4 is satisfied for typical expansion lines and working fluids, e.g.,

water and ammonia, heptane and toluene.

43
Theorem 1.4.2 (p-pinch for pumping configuration). Consider a regenerative Rank-

ine cycle without reheats and with positive isentropic efficiency of the turbine. Let

the turbine inlet temperature, inlet pressure, outlet temperature and outlet pressure

P be fixed, i.e., not influenced by the extraction. Consider an arbitraryfeedwater

heater, specified by an arbitrarybut fixed feed flowrate mF, feed inlet temperature TFi

and a heat transfer duty Q. Suppose that the feed pressure Pf is chosen so that the

feed stream remains subcooled. Suppose that the feedwater heater can be modeled as a

counter-current heat exchanger with a minimum approach temperature AMITAT and

without pressure drop. Assume that the extraction state is saturated or superheated,

i.e., the extraction temperature is not smaller than the saturation temperature of the

extraction pressure. Suppose that the drain is pumped to Pf and mixed with the feed

in the inlet or outlet of the FWH. Select a pair of extractionflowrate 7B and pressure

PB such that the cycle performance is optimized. Under Assumption 1.4.1 the MITA

occurs at the condensation onset. Moreover, subcooling of the drain to achieve the

double pinch adds heat transfer area without increase of efficiency.

Proof. We first show the necessity of pinch at the onset of condensation by contra-

position. Suppose that the pair (PB, MB) is optimal and the pinch does not occur at

the onset of condensation. For any pair (PB, rnB) the first law neglecting the pump

power gives

rnF,ihF,i+ iBhT(PB) = rhFohFo,

where the inlet state i is before mixing and the outlet state o after mixing. Similarly,

the mass balance gives

TfF,i + rnB = mpo-

Combining the last two equations we obtain

(T-F,o ~ fnB)hFi + rnBhT(PB) = r hF0

44
and therefore

rhB(hT(PB) - hFi)= - hFi)


,o(hFo

We differentiate and evaluate at (P, rnB)

(PB
(Or'B)
To -
(h,(PB) - hi)
M
B
0
PB PB
.OhT

By Assumption 1.4.1 we have hFi hT(PO) and thus

hT(PB) - hFi > hT(PB) - hT(P o )

and therefore

&WB
O PB
( &rnB~)
PB )To PB
(hT(PB) - hT(Po)) ± mB

M
hT
B PB
(111)

Similarly to the proof of Theorem 1.3.9 we have OWB > 0 and therefore PB is not

optimal.

We will now demonstrate that a double pinch is not advisable in terms of the

heat transfer area. Suppose that the pair (PB, MB) is optimal and the pinch occurs

at the onset of condensation and at the drain outlet. Keep the extraction conditions

and partition the heat exchanger into two segments: (i) for cooling the vapor and

condensation and (ii) for the subcooling. If we eliminate the second segment, we

still achieve complete condensation: in the case of pumping backward the inlets to

segment (i) are unchanged; in the case of pumping forward, the feed inlet to segment

(i) is colder resulting in higher heat transfer rate (the effect of lower flowrate on the

heat transfer rate is neglected). Moreover, eliminating the second segment does not

change the state of the feed outlet (after mixing); this is obvious by the first law and

mass balance. In conclusion, the heat transfer area can be reduced without loss in

45
performance. El

1.4.4 Open Feedwater Heater

For the sake of completeness, OFWHs are also considered. Clearly, OFWHs do not

fall in the same category as CFWHs and thus the double pinch criterion is not di-

rectly applicable. On the other hand, it is still worthwhile to optimize the extraction

pressure and flowrate. There are still three optimization variables for each OFWH,

namely the operating pressure and the bleed pressure and flowrate. Following the

same procedure as in the proof of Theorem 1.4.2, one can show that the optimal

extraction pressure is equal to the deaerator operating pressure. Moreover, the bleed

flowrate is given by the desired temperature increase and the requirement for satu-

ration at the feed outlet. Thus similarly to the CFWHs only one variable has to be

optimized for.

1.5 Numerical Examples with a Simple Flowsheet

In this section the validity of the design criterion is demonstrated for a single FWH

and multiple FWHs numerically. Power cycles with cascading and non-cascading

FWHs are considered. In addition to a prespecified minimum temperature approach,

the design criterion is demonstrated for the case of prespecified area. For the sake of

simplicity and compactness the same cycle is used to validate the criterion for both

single and multiple feedwater heaters.

A simple Rankine cycle implemented in AspenPlus® is shown in Figure 1-4 and

used to explain the importance of the double-pinch criterion. Feedwater exiting the

condenser, at the condenser pressure of 0.04bar and with a flowrate of 100kg/s, is

compressed to the boiler pressure 100bar, before entering the FWHs. Note that the

pressure in the condenser is below atmospheric which implies the need for a deaerator,

46
not modeled herein for simplicity; this does not affect the results and deaerator is

considered in Section 1.6. The temperature of the feedwater increases as thermal

energy is transferred from the bleeds passing through the heaters. Feedwater is then

heated in the boiler to a fixed outlet temperature of 500'C before entering the steam

expansion line where power is produced from the steam turbine. Two extractions,

one for each bleed, are present in the expansion line. Two bleed configurations are

shown, cascading and non-cascading. In both configurations, the bleed stream exiting

FWH2 is mixed with the main feedwater stream at the condenser. In the cascading

configuration (marked by x in Figure 1-4), the bleed at the exit of FWH1 is mixed

with the lower pressure bleed before entering the FWH2. In contrast, in the non-

cascading configuration (marked by o in Figure 1-4), the exiting bleed from FWH1

proceeds directly to the condenser. The cycle efficiency reported is the ratio of the

net produced power (turbine power minus pump power), to the heat transfer rate

in the boiler. For the sake of simplicity no pressure drops are considered and the

turbomachinery is assumed to be irreversible. Note that this nonrealistic assumption

does not affect the qualitative results and a more realistic case study is given in

Section 1.6.

1.5.1 Single Feedwater heater

Recall that the regeneration scheme, shown in Figure 1-4 has two closed FHWs. Only

the non-cascading configuration is discussed in this subsection and only FWH1 (high-

pressure) is analyzed. In contrast, the FWH2 (low-pressure) is considered fixed as

follows

PB2 = 0.158bar, rnB2 = 4.73kg/s, Q2 = 9.50MW

The specifications for FWH2 are chosen to result in a double pinch with a value of 3YC.

This arbitrary specification does not affect the results presented in this subsection,

47
Turbine
Boiler

Bleedi Bleed2

Condense
Cascading
FWH1

FWH2

N~on-Cascading Pump

Figure 1-4: Flowsheet for the numerical validation of the double pinch criterion.
Cascading bleed configuration marked by x and non-cascading bleed configuration
marked by o.

since they only affect the overall efficiency.

Minimal Approach Temperature

As aforementioned a common shortcut method in system-level analysis and opti-

mization is the pinch analysis. The proposed double-pinch criterion is validated nu-

merically for this shortcut calculation. The heat transfer duty in FWH1 is fixed to

Q, = 60.7MW. This value is selected based on approximately optimal performance

of the cycle. As aforementioned this heat transfer duty can be achieved for differ-

ent combinations of bleed flowrate and pressure, resulting in different MITA and heat

transfer area required. The bleed flowrate and pressure are discretized with 200 points

each, in the range

PB, E [13, 15]bar, rnB E [22, 24]kg/s

48
and the aforementioned flowsheet is simulated in AspenPlus@ for each value. The

results are illustrated in Figure 1-5, which shows contours of efficiency as a function

of the two variables. This corresponds to the optimization objective function and is

increasing with decreasing extraction pressure and flowrate. The figure also shows

the optimization constraints, namely the two possible pinch points, at the onset of

condensation and at the feedwater outlet for two given MITA. These pairs of lines

define the feasible region for higher pressure and/or flowrate (region to the upper

right), and the infeasible operation, i.e., operation violating a given MITA (region

to the lower left). The pairs of lines intersect at the double-pinch operation for a

given MITA and the figure shows the union of all intersection points. For a given

MITA the double pinch is at a higher contour line compared to either of the two

pinch lines. Mathematically, this can be expressed as the gradient of the objective

function lying in the feasible cone defined by the constraints. In other words if one

follows either of the pinch lines the efficiency increases towards the double pinch and

decreases away from it. From the graph two more facts are evident that can be also

proved analytically: (i) the pressure of the bleed at the double pinch is the smallest

extraction pressure that allows for a pinch at the outlet of the FWH; and (ii) the

flowrate of the bleed at the double pinch is the smallest flowrate that allows for a

pinch at the onset of phase change.

Maximizing the efficiency is equivalent to minimizing total entropy generation in

the system, or total exergy destruction. It is well-known that minimal system en-

tropy generation is not necessarily equivalent to minimal entropy generation for each

component. For instance, for a minimal entropy generation in the FWH a zero heat

transfer duty is preferable which results in a suboptimal cycle performance. However,

the numerical results suggest that for this simple flowsheet optimal design & oper-

ation of the FWH coincides with minimal entropy generation in the FWH for fixed

heat transfer duty and MITA. This is demonstrated in Figure 1-6 which plots the con-

49
24

23.9

23.8
-I p-pinch

0'
0.1 C C
c-
P

.
Efficiency
Double pinch

-p-pinch

o-pinch
s

23.7

23.6
-K feabibleI
23.5 0 ,
1 23.4 '.E
1
23.3 00 C Double pinch
0.1 s o-pinch
23.2 1% 2.5
23.1
Tlwtv....
H
13 13.5 14 14.5 15
Bleed extraction pressure (bar)

Figure 1-5: Contours of efficiency for regenerated duty Qi = 60.7MW for Flowsheet 1-
4. Pinch lines super imposed: pinch at onset of condensation with a MITA of 0.10 C
and 2.5 0 C (black line, labeled p-pinch); pinch at outlet of FWH with a MITA of
0.1 0 C and 2.5'C (purple line, labeled o-pinch); double-pinch with variable MITA
values (intersection of two pinch lines).
tours of entropy generation rate along with the aforementioned constraints. Minimal

entropy generation seems intuitively correct since the double-pinch seems to result in

smaller average temperature between the feed and bleed. However, as is evident from

Figure 1-1 (crossing of bleed lines) there is a tradeoff between extraction pressure

and flowrate, so that proving the validity of entropy generation would not be a trivial

task. Note also that minimizing for the entropy generation inside the cycle design is

not practical since it would require a constrained optimization problem, embedded

in the cycle simulation or optimization. For instance, the cycle optimization could

set/select the heat transfer duty and the entropy generation would be minimized by

varying the extraction pressure and flowrate subject to the MITA. Embedded opti-

mization problems are extremely challenging and only recently have they been solved

for nonconvex problems [14]. In other words minimizing entropy generation is deemed

more complicated than the original system-level optimization problem. The double

pinch approach on the other hand eliminates two degrees of freedom and satisfies the

design constraints at each iteration performed by the optimizer while eliminating the

need for a spatially distributed model. Finally, in the case of fixed heat transfer area,

minimal entropy generation in the FWH is not a good criterion as discussed in the

following.

Fixed Area

The results presented in the previous subsection validate the analytical proof derived

for the shortcut method of pinch analysis. This analysis in principle ignores capital

costs associated with increasing the heat transfer area by reaching the MITA in two

positions. To address capital costs, the flowsheet given in Figure 1-4 is now analyzed

for a given (constant) heat transfer area of FWH2, assumed equal to 2,516m 2 . To

demonstrate the generality of the results, this area is selected different than the

one corresponding to the selected heat transfer duty in the MITA. Similarly to the

51
24
p-pinch G gen
0.1
3.-
- p-pCh Double pinch
23.8 - 2. 5 S -
2.- p-pinch
23.7--
23.7-o-pinch

23.6- feasible -
23.5

23.4 -
0

23.3 Q Double pinch


apinch -p; c
C 'S-pinch
23.2

23.1

23
teas
- - 7
2.5 _

13 13.5 14 14.5 15
Bleed extraction Pressure (bar)

Figure 1-6: Contours of entropy generation rate $gen for regenerated duty Qi =
60.7MW for Flowsheet 1-4. Pinch lines super imposed: pinch at onset of condensation
with a MITA of 0.1'C and 2.5'C (black line, labeled p-pinch); pinch at outlet of FWH
with a MITA of 0.1'C and 2.5'C (purple line, labeled o-pinch); double-pinch with
variable MITA values (intersection of two pinch lines).
previous analysis the bleed flowrate and pressure are discretized with 200 points each,

in the range

PBI E [11, 13]bar, rhB1 E [21,23]kg/s.

This range is similar but not identical to before to account for a different heat transfer

area/duty. The results are shown in Figure 1-7 where the contours of efficiency are

plotted. The efficiency is maximal for middle values of the extraction pressure and

flowrate. For every extraction pressure there exists an extraction flowrate that max-

imizes cycle efficiency (green line); these pairs are near optimal. The difference is in

the order of 10 5 3
(10- percentage points), i.e., noticeable numerically but insignifi-

cant compared to model and/or numerical inaccuracies. In mathematical terms there

exists a linear manifold in the optimization variable space along which the directional

derivative is very small. In practical terms this allows the optimization of the cycle

even if the pressure cannot be selected with arbitrary accuracy. This is for instance

important for potential retrofit of existing cycles; therein it may not be possible to

change the extraction pressure but only the extraction flowrate. The small difference

in performance between the double-pinch pairs and the absolute optimum implies

that retrofiting may get almost the same performance increase as optimal design.

For low extraction pressures and high extraction flowrates (upper left corner) the

approach temperature at the onset of condensation is much smaller than the approach

temperature at the outlet and the opposite is true for high extraction pressures and

low extraction flowrates (lower right corner). Figure 1-7 also shows the line of (ap-

proximate) double pinch (red line), which results in an efficiency within 10-5 (10-3

percentage points) of the aforementioned near-optimal efficiency line. The difference

is so small that could be attributed to model/numerical inaccuracies and is not sig-

nificant from a practical perspective. It is also noteworthy that efficiency seems to

favor large pinch at the outlet versus large pinch at the onset of condensation. This is

a possible explanation for the current design practice of only slightly subcooling the

53
23
45.74 -45.79 4 45.89 45.92 45.97
22.8 22.846.011

22.6 region 460

22.4 -
46.03
22.2 -

22 + -- e . . 46.01

0 218 45.98

21.6

21.4 -p- 45.92

21
11 11.5 12 12.5 13
Bleed Extraction Pressure (bar)

Figure 1-7: Contours of efficiency for a fixed FWH Area 2,516 m3 . The green line
shows the optimal extraction flowrate for a given extraction pressure, while the red
curve shows the pairs that result in double-pinch; operation points on the two lines
give efficiency differences less than 10- 5 (10-3 percentage points). The blue cross
shows the optimal solution, while the black cross shows the best double-pinch point.

drain, see the following discussion. However, both unbalanced approach temperatures

are inferior to balanced MITA.

Figure 1-8 plots contours of entropy generation rate in FWH2 as a function of the

two bleed variables. The same lines as in Figure 1-7 are superimposed on the figure.

It is evident that minimal entropy generation in the FWH is not a good criterion for

maximal cycle efficiency. Recall that this is in contrast to the case of pinch analysis

for a given heat transfer duty. More concretely, minimal entropy generation occurs for

low heat transfer duty, which occurs at low extraction pressure and flowrate. In other

words, entropy generation minimization in the FWH ignores the benefits of increased

regeneration prior to the boiler. In contrast, the proposed double pinch criterion is a

good criterion for optimal efficiency.

54
23 146 149 150 151
22.8 -
p-pinch
22.6. region
22.4 max efficiency on
22.2 double pinch line
double pinch
22 - dblinh + ~ max efficiency
21.8

21.6

21.4

21.2
21
2- 141 141 142
11 11.5 12 12.5 13
Bleed Extraction Pressure (bar)

Figure 1-8: Contours of entropy generation gen for a fixed FWH Area 2,516 m
The green line shows the optimal extraction flowrate for a given extraction pressure,
while the red curve shows the pairs resulting in double-pinch; operation points on the
two lines give efficiency differences less than 10-5 (10-3 percentage points). The blue
cross shows the optimal solution, while the black cross shows the best double-pinch
point.
1.5.2 Non-Cascading, Cascading, and Common Practice

High-efficiency regenerative Rankine cycles have cascading bleeds in a FWH train,

i.e., combine the drain from a FWH with the inlet bleed of the preceding FWH (next

lower pressure). In Figure 1-4 the outlet from FWH1 is mixed with the inlet to FWH2

(line marked by x). The advantage of this arrangement is that the outlet bleed still

has significantly higher temperature than the following deaerator or the condenser

and thus the availability of the stream can be used to preheat the feedwater and thus

reduce the required bleed flowrate to the preceding FWH. Typically, the cascading

FWH are designed and operated to achieve the MITA in the onset of condensation

and subcool the outlet bleed by a few K. This seemingly reduces the heat transfer area

needed without loss in performance, since the bleed will be further used. However,

this analysis may be misleading since further subcooling the bleed would imply that

the preceding FWH needs to preheat the feed to a lower temperature.

Recall that no analytical proof for the optimality of cascading double pinch is

given herein. Instead, the criterion is examined numerically for the flowsheet given

in Figure 1-4. Comparing the efficiency with the pinch analysis approach could be

seen as unfairly favoring the proposed double-pinch criterion due to potentially larger

heat transfer area. Consequently, the comparison is done for a constant total heat

transfer area. The following four configurations/designs are compared: (i) cascading

configuration with the proposed double-pinch criterion; (ii) noncascading configura-

tion with the proposed double-pinch criterion; (iii) cascading configuration with the

current design practice of slight subcooling at outlet for the FWH1 and the proposed

double-pinch criterion for FWH2 (low pressure); (iv) cascading configuration with

the current design practice of slight subcooling at outlet for both FWHs. For each

case, a cycle-level optimization of the efficiency is performed by varying the fraction

of heat transfer area between the two FWHs as well as the extraction pressures. For

the double-pinch criterion, for a given pressure and heat transfer area, the FWH is

56
fully specified, by (1.10). In the current design practice, for a given operating pres-

sure the bleed outlet temperature is specified as Tt(PB) - 2K; therefore, for a given

heat transfer area the FWH is fully specified. A simple thought experiment to verify

this is to note that the inlet temperature of the feed is fixed, and so is the inlet and

the outlet temperature of the bleed; the feed flowrate is also given, so if we select

the bleed flowrate we obtain the heat transfer duty and feed outlet temperature by

energy balance; heat transfer correlations result in calculating the heat transfer area.

In AspenPlus@ the bleed flowrates and heat transfer duty for each FWH are imple-

mented by design specifications embedded into the optimization. For a motivation

for this decomposition, see [24].

The calculations are performed for heat transfer coefficients accounting for the dif-

ferent regimes, i.e., for the vapor-fluid section U = 0.709kW/(m 2 K), for the conden-

sation section U = 3.975kW/(m 2 K), and subcooling section U = 1.704kW/(m 2 K), as

taken from an example in [15]. Note that the values of the heat transfer coefficients

are actually dependent on the heater's geometry and flow conditions but here are

taken as constant for simplicity. Additionally, calculations for a constant overall heat

transfer coefficient are performed, but since these result in very similar qualitatively

results, they are not shown for the sake of compactness.

Figure 1-9 plots the optimal efficiency for each of the four design procedures as

a function of the total area of the FWHs. As expected all four curves are monoton-

ically increasing indicating the tradeoff of capital costs and efficiency, including the

asymptote to a finite value for infinite heat transfer area. Moreover, as expected,

the cascading flowsheet with double pinch outperforms the noncascading equivalent.

The main finding is that the cascading cycle with double pinch in both feedwater

heaters outperforms significantly the standard practice. For large values of the heat

transfer area, the efficiency improvement is in the order of 2 percentage points com-

pared to slight subcooling in both FWHs and in the order of 0.5 percentage points

57
47 r

46.8-

Co
46.4
dp cas
T e - practice pinch
o 46. - - -N-dp non
F9p- practice cas
.E 46 -
O ~
45.8

45.6
500 1000 1500 2000 2500 3000 3500 4000
Total Regeneration Area (MPf)

Figure 1-9: Optimal performance of four different design procedures versus total
FWHs area for flowsheet given in Figure 1-4: cascading with double-pinch for both
FWHs (green solid curve, labeled "2pinch-2pinch cascading"); cascading with sub-
cooling of 2K for FWH1 and double-pinch for FWH2 (purple dashed-doted line, la-
beled "ppinch-2pinch cascading"); noncascading with double-pinch for both FWHs
(blue line with x marks, labeled "2pinch-2pinch noncascading"); cascading with sub-
cooling of 2K for both FWHs (red line with triangle marks, labeled "ppinch-ppinch
cascading").

for the case that the double pinch criterion is applied to the low-pressure FWH and

slight subcooling is applied to the high-pressure FWH. For small values of the heat

transfer area, the efficiency improvements are not as dramatic but still substantial.

Moreover, the noncascading cycle with a double pinch outperforms the cascading cy-

cle without double pinches, and is very close to the cascading configuration with a

double pinch in FWH2. Finally, the difference in performance among the different

procedures increases with increasing area, or the proposed design criterion becomes

more important for large heat exchanger areas.

58
1.6 Numerical Case Study with a Realistic Cycle

Design

In this section a realistic Rankine cycle is considered, illustrated in Figure 1-10. It

contains four CFWHs and an OFWH acting as the deaerator. The four CFWHs

are arranged in two pairs, above and below the deaerator. For each of the two

pairs of CFWHs cascading is used, i.e., the drain of the high temperature CFWH

is combined with the bleed entering the CFWH. The drain of the lowest pressure

is pumped upwards. The cycle specification are shown in Table 1.1. For simplicity

the expansion line is considered to have a constant isentropic efficiency. Otherwise,

the optimization is significantly complicated, see the discussion on integer variables

in [24].

The proposed optimization criterion is compared with the current design practice

of small subcooling in the drain. Initially, the bleeds are optimized following a MITA

specification of 2K for each CFWH (FWH1,2,4&5) and with an subcooling of the

drain of 2K. Then the area of each CFWH is fixed and used for optimization of the

flowsheet with the proposed double-pinch approach for three of the four CFWHs.

Since the drain of the last CFWH is pumped upward, the double-pinch is not optimal

and the drain is subcooled by 2K. Moreover, the bleed is two-phase liquid (not

superheated) and thus a pinch occurs at the inlet of the bleed. The results are shown

in Table 1.2; the proposed criterion results in a significant efficiency increase, in the

order of 0.45 percentage points. Note that this is achieved merely by changing the

bleed pressures and flowrates, without any increase in heat transfer area, without

addition of components, and without changing the flowsheet connectivity. Moreover,

the area of each FWH is selected based on optimization of the conventional design

criterion; allowing for a redistribution of the heat transfer area would result in further

savings for the proposed criterion.

59
Turbine
Boiler--

BleedB Bleed3
Bleed2 Bleed4 Bleed5

FWH1

FWH4 Condes

FWH2H

FWH2 ~ Deaerator
F H
FWH(p) LP Pump

Bleed Pump
HP Pump

Figure 1-10: Realistic cycle design with four closed feedwater heaters

Table 1.1: Specifications of flowsheet with 4+1 FWHs in Figure 1-10

Unit Name Unit specifications


Feedwater main flowrate 108 kg/s
Boiler pressure 150.3 bar
Boiler superheat temperature 542.9 0 C
Turbine efficiency Isentropic 0.7 - Mechanical 1
Deaerator pressure=bleed3 pressure 17.53 bar
Condenser pressure 0.05 bar
Condensate temperature 33.0 0 C
Pumps Efficiency Isentropic 1 - Mechanical 1
LP Pump discharge pressure 20.31 bar
Specification Conventional Double-pinch
FWH1 MITA = 20 C Area = 2165 m 2
FWH2 MITA = 2'C Area = 1667.8 m 2
FWH4 MITA = 20 C Area = 1486.8 m 2
FWH5 MITA = 2'C Area = 1048 m 2
Table 1.2: Results of flowsheet with 4+1 FWHs in Figure 1-10

Optimization results Conventional Double-pinch

Efficiencyqr38.11% 38.56 %
PM' 71.3 bar 82.87 bar

rm) 14.9 kg/s 15.81 kg/s

p(2 ) 36.0 bar 35.21 bar

7h(2 6.26 kg/s 6.736 kg/s

7h(3 6.221 kg/s 8.418 kg/s

p(4 ) 5.883 bar 4.619 bar

rh() 9.120 kg/s 8.212 kg/s

p(5) 1.080 bar 0.7478 bar

mh(5 7.162 kg/s 6.621 kg/s


1.7 Conclusion and Future Work

A new design criterion is proposed for the design and operation of feedwater heaters

in regenerative Rankine cycles. The basis is to have the same pinch in the onset

of condensation of the bleed and in the outlet of the bleed. The criterion is proved

analytically for a simple configuration and illustrated numerically in case studies for

various configuration, see Table 1.3. Application of the criterion results in signifi-

cant efficiency improvements for a constant heat transfer area (representing capital

costs). Moreover, a procedure is proposed that drastically simplifies the design and

optimization of regenerative Rankine cycles. In the pinch analysis, for each feed-

water heater the pinch value and extraction pressure (design variable), are fixed or

optimized for; the bleed flowrate and heat transfer rate (operational variables) are

adjusted to achieve the double pinch. In the rigorous calculation, the extraction pres-

sure and heat transfer area (design variables) are fixed or optimized for; the bleed

flowrate and pinch value are adjusted to achieve the double pinch. The case studies

demonstrate that under the proposed double-pinch criterion, the cycle performance is

not very sensitive to the design and substantial improvements to performance can be

achieved by adjusting only the operational variables. If local solvers are used for the

optimization, the criterion increases the chances to find a global optimum; if global

solvers are used the number of variabls and constraints is reduced which typically re-

sults in significantly faster CPU times. Regenerative Rankine cycles are very common

in industry and novel energy systems and thus the presented criterion has important

implications for research & development.

Future work should include experimental validation for both existing and new

cycles. Additionally, consideration of controlability of the proposed operation and

second law analysis is of interest. Moreover, the double pinch criterion could be

applied to different systems, such as boilers and heat recovery steam generators, and

cases where both streams exhibit phase change. Finally, it would be interesting to

62
consider splitting the drain and use in both cascading and non-cascading way.

Table 1.3: The applicability of the proposed design criterion for various configurations
and the evidence given in this chapter.

Pinch analysis with fixed Fixed heat transfer area


Q, and MITA
Drain to condenser double pinch unique opti- double pinch unique op-
mum (analytical proof + timum (numerical case
numerical case studies) studies)
Drain to OFWH double pinch unique op- double pinch unique op-
timum (numerical case timum (numerical case
studies) studies)
Cascading drain to double pinch unique op- double pinch unique op-
CFWH timum (numerical case timum (numerical case
studies) studies)
Pumping back- double pinch non-unique double pinch not optimal
ward/forward (down- optimum (analytical (analytical proof)
wards/upwards) proof)
OFWH double pinch not appli- double pinch not appli-
cable; optimal extraction cable; optimal extraction
pressure equals to oper- pressure equals to oper-
ating pressure (analytical ating pressure (analytical
proof) proof)

63
64
Chapter 2

Optimal Design and Operation of

Pressurized Oxy-Coal Combustion

with a Direct Contact Separation

Column

2.1 Summary

Simultaneous multi-variable gradient-based optimization is performed on a 300 MWe

wet-recycling pressurized oxy-coal combustion process with carbon capture and se-

questration. A direct contact separation column is utilized for practical and reliable

low-temperature thermal recovery. The models for the components include realis-

tic behavior like heat losses, steam leaks, pressure drops, and cycle irreversibilities.

Moreover, constraints are used for technoeconomical considerations. Optimization

involves 17 optimization variables and 10 constraints, with the objective of maximiz-

ing the thermal efficiency. The optimization procedure utilizes recent design rules

and optimization procedures for optimal Rankine cycle performance, as explained in

65
Chapter 1, speeding up the plant optimization process by eliminating variables and

avoiding constraint violations. Moreover, the procedure partially alleviates conver-

gence to suboptimal local optima. The basecase of the study is a comprehensively

optimized cycle that utilizes a surface heat exchanger, a more thermodynamically-

effective form of thermal recovery which however bears significant materials chal-

lenges. Upon optimization, the cycle utilizing the direct column is seen to be very

attractive regarding efficiency and performance. Moreover, the optimization results

unveil potential for reducing capital costs by eliminating the first carbon sequestration

intercooled compressor and by showing possibilities of process intensification between

the separation column and the carbon sequestration purification columns.

2.2 Motivation

The importance of emissions free power generation is motivated and discussed ex-

tensively in literature [16, 17]. Clean and renewable power production are of high

interest to both academic and industrial research aiming to make such technologies

more affordable and reliable. However, the world's dependence on fossil fuels for power

generation, especially coal due to its cheap price and abundance of reserves [181, is

expected to continue at least till renewable power generation becomes more econom-

ically attractive.

Pressurized Oxy-Coal Combustion (OCC) with Carbon Capture and Sequestra-

tion (CCS) mitigates the emissions problem while relying on the cheapest fossil fuel

[19, 20, 21, 22, 23, 24]. In OCC the flue gas is mainly carbon dioxide and water vapor,

and the latter can be separated by condensation. Flue gas cooling and condensation

can be integrated to recover thermal energy, particularly latent, into the low tempera-

ture section of the power cycle, [20, 21, 23, 24, 25]. As flue gas pressure increases, the

vapor dew point in the flue gas increases allowing for condensation to occur earlier

66
and at a higher temperature. This increases the amount of recovered latent energy

and increases its quality since it occurs at a higher temperature. Pressurizing the

combustion process increases the compression requirements of the air separation and

oxygen delivery process while reducing those for the carbon sequestration process,

but also contributes in increasing the pressure losses and irreversibilities within the

flue gas; the tradeoffs signify a presence of an optimum operation. Simultaneous

multi-variable optimization, like the one dealt with in [24], is required to obtain the

optimum operation and achieve an attractive cycle performance. Optimization in [24]

contributes in significant efficiency increase, 0.76% points over the literature proposal

of 10bar combustor pressure, [211, while simultaneously reducing the combustor's op-

erating pressure, to the range of 7.41 bar, thus making the process more attractive

and practical. Efficiency is 3.12% points higher than that of the atmospheric oper-

ation. Results also show the importance of the 15 other optimization variables in

obtaining such efficiency improvements.

The pressurized OCC cycle presented in [24] utilizes a recovery heat exchanger

(RHE), which is a surface heat exchanger, for recovering thermal energy from the

water vapor present in the flue gas. However, this type of heat exchanger is subjected

to considerable amount of fouling and damage from the contaminated flue gas. While

the surface heat exchanger is thermodynamically more efficient, it requires relatively

more care and maintenance. A more practical form of thermal heat recovery which is

less susceptible to fouling is a direct contact separation column (DCSC). Comparison

of the capital and operating costs between the two recovery units is out of the scope

of this work.

Separation columns are used in various engineering and chemical processes [26],

and in fact are used as part of the Carbon Sequestration Unit (CSU) in the OCC

cycle, where the dry flue gas is purified from nitrogen and sulfur oxides and other

contaminants. In general the separation process is performed by having two streams

67
in a vertical counter flow arrangement where undesired substances in one stream are

transferred to the other stream. In this study, a DCSC is used instead of a surface

heat exchanger to condense water vapor from the flue gas and recover some of the

latent and sensible energy.

Replacing the RHE with a DCSC changes the cycle's performance substantially.

The DCSC utilizes an intermediate stream between the flue gas and the working

fluid of the power cycle for the recovery process, leading to a less effective thermal

recovery compared to a RHE. Therefore, the efficiency of the cycle with the DCSC

is expected to decrease. The difference in the operation and performance of the two

units mandates optimization of the operating conditions for the DCSC flowsheet,

which are expected to be noticeably different than those for the RHE flowsheet.

Herein, multi-variable gradient based optimization is performed for the model

presented in [27, 24] with a DCSC replacing the RHE. A similar methodology and

approach are followed as those explained in [27, 24], the optimization results of which

are taken as the basecase of the current work. Recent design rules and optimization

procedures, [27, 24, 28, 29] and Chapter 1, are incorporated and automated within

the model. Detailed and high fidelity modeling of components and irreversibilities

are also considered to accurately assess the advantages and tradeoffs compared to the

original RHE flowsheet and compared to other coal CCS technologies. The details of

the model and the specifications are presented in Section 2.3. Section 2.4 describes

the DCSC unit and presents its modeling approach and simulation analysis for the

proper integration within the pressurized OCC flowsheet. Section 2.5 deals with the

optimization formulation and describes the objective function, optimization variables,

and optimization constraints. Results are shown in Section 2.6 where the influence

of the critical variables on the cycle are analyzed. Results also suggest possibilities of

capital cost reductions.

68
2.3 Flowsheet and Model Description

2.3.1 Power Plant Flowsheet

The flowsheet and model specifications studied here are identical to those of [24]

with the exception of utilizing a DCSC instead of a RHE. Aspen Plus@ is also

used for modeling the work presented here. Figure 2-1 shows the schematic of the

flowsheet with the different sections, Air Separation Unit (ASU), Combustor, Rankine

Cycle, DCSC, and CSU. Oxygen is separated from air by the ASU and provided at

an elevated pressure to the combustor as an oxidizer. The combustor is based on

the ISOTHERM PWR® technology, [30], patented by ITEA [31, 32, 33]. Prior to

combustion, the pressurized oxygen is mixed with the primary recycled flue gas stream

(FG-Rec-pri) to control the combustion to a temperature to 1550'C. Combustion

gas (Comb-Gas) exiting the Combustor is mixed with a secondary recycling stream

(FG-Rec-sec), forming Hot-Gas, before entering the Heat Recovery Steam Generator

(HRSG) to maintain a temperature of 800 0 C specified by the metallurgic properties

of the HRSG. The HRSG is based on a proprietary ITEA/Ansaldo Caldaie design,

developed with the support of ENEL. The HRSG is the site of main thermal energy

transfer from the flue gas to the Rankine cycle working fluid. Upstream of the DCSC,

the flue gas temperature should remain above the acid condensation temperature.

Sulfur and nitric oxides resulting from the combustion of coal cause damage to the

material and components if they condense outside the DCSC. Therefore, constraints

on Cool-Gas and on the feedwater entering the HRSG, FW-HRSG-in, are placed with

safety margins to avoid condensation in the flue gas and on the tubes of the feedwater

respectively. A large fraction of the Cool-Gas exiting the HRSG is recirculated for the

temperature control processes utilizing fans that compensate for the flue gas pressure

drops. The flue gas pressure losses occur mainly in the HRSG and the recirculating

pipes. The non recirculated fraction of Cool-Gas proceeds to the DCSC where acids

69
are condensed from the flue gas before the CCS process. Moreover, low quality

thermal energy, approximately 40'C to 220'C latent and up to 300'C sensible, is

recovered by cooling the flue gas and condensing the water vapor. FG-DCSC-out exits

the DCSC and proceeds to the CSU where it is further purified, compressed to 80bar,

liquefied and pumped to 110 bar. The high pressures used allow the liquefaction at

the ambient temperature eliminating the expensive operation and capital cost that

would otherwise be needed for cooling and storage.

The power cycle utilized is a supercritical, single reheat, regenerative Rankine

cycle. Only high-pressure feedwater heaters (FWHs) are utilized for regeneration;

the thermal recovery from the DCSC and the thermal energy absorbed from the

combustor's losses reduce the benefits of adding low-pressure FWHs preceding the

deaerator. The steam expansion line is accurately represented by twelve stages with

specified isentropic efficiency for each stage, [21, 24]. Four extractions are required

from the expansion line, one for each bleed (two closed FWHs and one open FWH

or deaerator) and one for the combustor's atomizer stream. A detailed description of

the flowsheet and model is found in [24]. Table 2.1 summarizes the flowsheet's fixed

parameters.

2.3.2 Flue Gas Pressure Losses

For the accurate assessment of the cycle and in order to find realistic optimum oper-

ating conditions, losses and irreversibilities have to be accounted for. As detailed in

[24], the pressure losses in the recycling pipes are calculated as:

APp = pf (2.1)

70
Table 2.1: Fixed Simulation Parameters. The superheat pressure is lower than the
highest pressure in the cycle due to pressure losses through feedwater heaters, con-
nection pipes, and the HRSG.

Simulation-Parameter Parameter Value


Name
Coal HHV 31.09MJ/kg
Coal LHV 29.88MJ/kg
Coal mass flowrate 30kg/s
Slurry water input mass 35.48%
fraction
Atomizer Stream weight ra- 8.33%
tio (steam/Coal)
Atomizer stream pressure 30 bar
Oxygen purity 95%
Oxygen molar fraction in 3%
flue gas
ASU specific work for 95% 837kJ/kg
purity at 1.238bar
SuperHeat stream pressure 250bar
SuperHeat stream tempera- 600 0C
ture
Reheat stream pressure 55.67bar
Reheat stream temperature 610 0C
Condenser operating pres- 0.04bar
sure
Condenser operating tem- 29.7 0C
perature
HP-Pump pressure (maxi- 266.2bar
mum water pressure)
ASU
Pressurized
02 02
Compressor r

02 02
sep
N2
a-

-G-Rec-pri FG-Rec-se tW-HRSG-in Reheat

HRSG CSU

CoalT Wati CombustioSe estrated C2


Slurry .. Gs ot-G i o-Gas -C -o
.... Te ure FG-DCSC -in---.....
vented gas
FW-DCSC-in
Ash Controller

Atomizer
Condensate

H P IPT LPT

Bleed Bleed Bleed3

Reheat Cooling Water

FW-HRSG-in FWHD FWH Deaerator Condenser

HP-Pump

LP-Pump

Figure 2-1: Oxycombustion cycle flowsheet based on wet recycling utilizing a DCSC.
Note that this schematic does not represent entirely the modeling, e.g., turbines
were modeled with multiple stages in Aspen Plus. The DCSC flowsheet is shown in
Figure 2-2
where V is the bulk gas velocity in the pipe, d is the pipe diameter, Lp is the pipe

equivalent length, p is the gas density, and f is the friction factor calculated by:

fpe = "1 -2.0log


2 l 2/d)_
[7.4
. 2 log
Red 7.4
+ 13
Red
-

}
-2

where e is the pipe roughness, Red = PVd is the Reynolds number based on the pipe

diameter, p is the flue gas density, and p is the dynamic viscosity of the gas.

The HRSG pressure drop is calculated as:

APHRSG,a - QHRSG,aPaTib (2.2)


APHRSG,b QHRSG,bp~rha

where subscripts a and b stands for actual and the basecase of the initial RHE

flowsheet optimization respectively. QHRSG is the thermal energy transferred in the

HRSG, and mh is the flowrate of the flue gas. A detailed derivation of the equation is

provided in [24].

2.4 DCSC Modeling

2.4.1 DCSC Flowsheet

The DCSC unit is shown in Figure 2-2. Flue gas enters the bottom of the eight

stage separation column and exits from the top stage. Eight stages are selected to

achieve acceptable separation; a detailed design of the column is outside the scope

of this work. Meanwhile, a recirculating water stream, Rec-Wtr-Sep-in, flowing in

the opposite direction, from top to bottom, cools down the flue gas and absorbs the

condensate. The condensate is mainly water vapor along with sulfuric and nitric

acids; despite the sulfuric and nitric acids condensation in the separation column, the

73
flue gas still requires purification in the CSU to fit sequestration standards. Although

the acid condensation does not result in high acidity in the circulation water as it

passes through the separation column, NaOH is introduced gradually into the stages

of the different stages to maintain a neutral pH. The amount of NaOH needed is

very small compared to the large flowrate of the condensed water and much larger

flowrate of the recirculation water, and thus does not affect the mass and energy

balance. Also, the concentration of salt is much smaller than the solubility limit

and thus no precipitation occurs. The recirculation water stream itself heats up and

increases in mass flowrate, as it collects the condensing water, forming RW-Sep-out.

Flue gas exits the separation column with a reduced temperature and vapor fraction

depending on the flue gas pressure and the recirculation water stream temperature

and flowrate. The separation column is modeled by a RadFrac column in Aspen Plus

with no condenser nor reboiler.

The ability of the DCSC to cool and dehumidify the flue gas is expected to be

less than that of a RHE. In more details, the temperature of the recycling stream

in the DCSC recovery unit, the cold stream in the separator, is always greater than

the temperature of the working fluid of the Rankine cycle exiting the low-pressure

pump, which is the cold stream of the RHE in the RHE cycle. This is because the

recirculating water and the Rankine cycle working fluid exchange thermal energy in

the DCSC-HX which has a positive minimum internal temperature approach (MITA)

specification as explained later. This in principle causes smaller temperature decrease

of the flue gas and smaller condensation rates for the DCSC compared to the RHE.

Moreover, in the separation column, the condensed water increases the capacity of

the cold stream causing a smaller temperature rise of the recirculating water.

As a result of the expected higher temperature and larger humidity of the flue gas,

a flue gas cooler interacting with the atmosphere is introduced after the separation

column. The cooler is allowed to reduce the flue gas temperature to 36.9'C, which

74
7FG-Sep-out FG-DCSC-out
Flue Gas Cooler/ Chiller Condensed Water
ChiN Condensate
_RW-Sep-in

Separation Clp (nlow


Column ain t RW-HX-out/RW-Split-in

FW-HX-outS-H
FG-Sep-in

-. I FW-HX-in

RW-Sep-out
R-Xi

PUMP

Figure 2-2: Direct Contact Separation Column (DCSC) operation unit for the low
quality thermal energy recovery

is the minimum flue gas temperature obtained in the RHE flowsheet, [241, without

introducing a cooler. This temperature is obtained based on the MITA on the RHE

(7.1'C [24]) and the temperature of the water in the Rankine cycle at the exit of the

low-pressure pump. The temperature at the exit of the low-pressure pump is defined

by the condenser's fixed operating pressure and the specifications of the pump itself.

Cooling close to atmospheric temperature is achievable and attractive because it

reduces the CSU compression power requirements. Moreover, the chiller prevents

unacceptably large humidity in the flue gas entering the CSU. Since the air and the

coal flowrates are fixed input parameters, the flowrate of the flue gas entering the

CSU changes only due to a change in humidity, which is relatively small. Thus,

the flue gas pressure and its pressure losses are the dominant factors in the CSU

power requirements. The chiller temperature is limited to 36.9*C because a lower

temperature requires a larger, more expensive, cooler to drive the thermal energy

across a low temperature gradient, and because no further cooling across the same

temperature gradient is considered for the RHE flowsheet.

The recirculating water stream exits the separator column from the bottom stage

75
carrying the excess water and flows through a pump that compensates for any pressure

losses. Then recirculating water enters the heat exchanger (DCSC-HX) where thermal

energy is transferred to the working fluid of the Rankine cycle. DCSC-HX is defined

by a MITA of 5'C. The recirculating water stream exits the DCSC-HX where a splitter

is responsible for controlling the flowrate back to its original value and closing the

recirculating water loop while rejecting the Excess-stream. DCSC-HX is introduced

before the splitter in order to allow for a larger thermal energy transfer from the

recirculating water stream and the carried condensate as opposed to the recirculating

water stream only. In other words, the excess stream separated by the splitter exits

the loop at a lower temperature when the splitter is after the DCSC-HX as opposed

to being before.

No pressure losses are considered in the DCSC since losses in this section are minor;

the flue gas operating pressure is expected to be relatively large, so the pressure loss

within the separator column which is of the order of 0.lbar is not significant compared

to an operating pressure of the order of 10bar, and thus barely affects the CSU power

requirements. Moreover, the recirculating water stream is a liquid, so the pump power

requirements required to compensate for the liquid losses are also insignificant. Note

that in the RHE flowsheet [24], the thermal recovery did not include pressure losses

and therefore it is consistent to do the same for the DCSC.

For all units in the DCSC flowsheet the Electrolyte NRTL property method is

used. The DCSC-HX handles the Rankine cycle working fluid (cold side of the heat

exchanger) with the Steam Tables physical property. A set of reactions describing

the nitrogen and sulfur oxides formation and reactions are used in the Separation

Column and are given in Appendix A.

76
2.4.2 DCSC Operation

A detailed design and sizing of the DCSC requires thorough modeling of phase and

chemical equilibrium and possibly kinetic and residence time. However, herein only

the aspects that are relevant to the power plant operation need to be considered. By

evaluating extreme cases of species' chemical reactions, it is found that kinetics in

the separator column play a minor role on the overall energy and mass balance. The

extreme cases of flue gas with no S0 2 /SO 3 conversion versus a complete conversion

has minor effects on the temperature of the involved streams and thus insignificant

effect on the amount of thermal energy recovered into the Rankine cycle. Therefore,

kinetics and separator column sizing are not incorporated in the integrated flowsheet

of the total pressurized OCC cycle.

The main variables influencing the DCSC and the amount of thermal energy re-

covered are the flue gas pressure, specifying the separation column operating pressure,

and the characteristics of the recirculating water stream. Due to the active constraint

optimization approach, explained in Section 2.5.5, the recirculating water character-

istics are not independent. The stream pressure is equal to the flue gas/separation

column pressure since pressure losses in the unit are considered minor and neglected.

Any other specification of the recirculating water stream, e.g. the recirculating wa-

ter temperature or a species concentration is enough to define the DCSC behavior.

Changes to the flowrate of the recirculated water result in the easiest convergence of

the stream's specifications within the recirculation loop and thus considered as the

optimization variable associated with the DCSC. The recirculating water flowrate is

also the simplest characteristic to measure and implement during actual operation.

Section 2.6.1 explains how the stream flowrate specifies the state of the recirculating

water.

The operating pressure of the DCSC flowsheet is expected to be larger than that

of the RHE flowsheet in order to allow for enough water condensation and thermal

77
recovery. The flowrate of the recirculating water, which dictates the stream's own

temperature, is another important variable in thermal recovery. It also determines

the flue gas temperature at the exit of the separation column where the chiller is

initially seen important in order to lower the CSU power requirements.

2.5 Optimization Formulation

The importance of simultaneous multi-variable optimization is illustrated in [24],

where significant improvements are obtained compared to a single variable sensitivity

analysis. Therefore, for the DCSC pressurized OCC cycle a similar methodology is
applied. The basecase of the DCSC model is the optimum operation of the RHE

model. Optimization is performed within Aspen Plus using the built-in SQP op-

timizer. Multi-start, similar to that in [24, 27], is performed here to increase the

probability of finding the global optimum, and disregard suboptimal local solutions.

2.5.1 Objective Fanction

The objective considered is to maximize the thermal efficiency of the cycle. The

fuel flowrate and specifications are fixed, therefore, the objective is equivalent to

maximizing the net power output. The net power is equal to the total power produced

from the Rankine cycle minus the power consumption of the pumps, the recirculating

fans, the ASU, and the CSU. Economical concerns are accounted for by considering

reasonable HRSG size, recycling pipe diameters, MITA for heat exchangers, etc.,

similar to the procedure in [24].

2.5.2 Optimization Variables and Constraints

Implementing the DCSC instead of the RHE adds to the complexity of the cycle

on both levels, simulation and optimization. At the simulation level, convergence

78
is required for the separator column and the recirculating stream. More critically,

at the optimization level, the unit adds to the variable count. The RHE flowsheet

model consists of 10 constraints and 13 optimization variables, and an additional three

integer variables are required for the accurate representation of the steam expansion

line. The DCSC flowsheet model incorporates 10 constraints and 14 optimization

variables, in addition to the three integer variables. As discussed in Section 2.5.5,

some constraints and variables were coupled in design specifications to accelerate

convergence, avoid constraints' violations, and avoid suboptimal solutions. Figure 2-3

shows the optimization variables in purple circles marked by (o) and the optimization

constraints in green circles marked by (x).

Optimization Variables

The DCSC flowsheet has 15 optimization variables in common with the RHE flow-

sheet, including the three integer variables required for defining the bleeds extraction

stages. The variable initially associated with the RHE, TRHE, the temperature of flue

gas at the exit of the RHE which specifies the amount of thermal recovery in the

RHE, is replaced with the DCSC variables: (i) thermal energy transferred in DCSC-

HX, QDCSC-HX, and (ii) recirculation water flowrate, 72RW-SePin. These two variables

essentially determine how much thermal energy is transferred from the flue gas into

the recirculating water stream and eventually into the working fluid of the Rankine

cycle. Table 2.2 shows the optimization variables as they appear in Figure 2-3. The

optimization variables are all independent, for example the bleeds' extraction pres-

sures, PBLD, and flowrates, rnBLD, can be manipulated separately. In contrast, the

temperature of the bleed is not independent of the extraction pressure and therefore

not considered as a variable. Similarly, the duty transfer in the feedwater heaters,

QFWH, are independent variables that define the regeneration from the bleeds to the
feedwater. The recirculating water flowrate in the DCSC, nRec-Sep-n, and the duty

79
transfer in the DCSC-HX, QDCSC-HX, are optimization variables but not the temper-

ature of the recirculating water because it is defined by the recirculation flowrate, the

flue gas conditions entering the separation column especially the flue gas pressure,

and the amount of thermal energy transferred in the DCSC-HX.

Table 2.2: Optimization Variables. The deaerator operating pressure should be above
atmospheric, but here the lower bound was intentionally taken sub-atmospheric to
examine if it would lead to any advantage in performance.

Number Variable Range Base-case default value


1 PComb [1.283 - 30] bar 7.41 bar
2 QComb [1 - 30] MW 1 MW
3 mFW,Main [240-340] kg/s 306 kg/s
4 PBLD1 [30 - 250] bar 99.0 bar
5 rnBLD1 [0 60] kg/s
- 62.2 kg/s
6 PBLD2 [10 - 120] bar 26.0 bar
7 rnBLD2 [0 - 30] kg/s 14.7 kg/s
8 PBLD3 [4.5 - 30] bar 16.8 bar
9 rnBLD3 [0 - 30] kg/s 9.27 kg/s
10 QFWH1 [0 - 200] MW 138 MW
11 QFWH2 [0 - 200] MW 37.7 MW
12 PDeaerator [0.1 - 30] bar 14.7 bar
13 QDCSC-HX [20 - 130] MW 44.9 MW
14 maec-sep-in [50 - 500] kg/s 100 kg/s
15 BLDILstage integer variable Stages: 1-4 Stage: 3
16 BLD2-stage integer variable Stages: 3-6 Stage: 5
17 BLD3_stage integer variable Stages: 5-7 Stage: 6

Optimization Constraints

Constraints define the allowable limits of operation. The limits are dictated by phys-

ical, practical, or economical considerations. Nine optimization constraints in the

DCSC flowsheet are identical to those in the RHE flowsheet. The MITA constraint

on DCSC-HX replaces the MITA constraint on the RHE. Table 2.3 states the DCSC

constraints as they appear in Figure 2-3.

80
Table 2.3: Optimization Constraints for the pressurized OCC process
utilizing a DCSC.
Value
Nu mber Constraint Value
1 MITAHRSG 3.70 C a
2 MITAFWH1 2.1 0C
3 MITAFWH2 2.1 0 C
4 MITADCSC-HX 5.0 0 C
5 qDeaerator Saturated Liquid
6 Tcool-Gas 20 0 C above acid condensa-
tion temperature
7 TFW-HRSG-in 5YC above acid condensa-
tion temperature
8 Tcomb-Gas-in 20 0 C above acid condensa-
tion temperature
9 CO 2 -Cap 94% of total CO 2 produced
10 CO 2 -Pur 96.5% purity

a The temperature of the flue gas is initially high as it enters the HRSG, 800'C;
it is guaranteed that the temperature difference between the flue gas and the
water/stream of the Rankine cycle, which reach a maximum of 610'C, is large
in the superheat of the main feedwater and the reheat sections; the
temperature approach in the lower temperature sections of the HRSG is
limiting due to the flue gas temperature drop. A temperature approach
around 4C is common, [34]

2.5.3 Integer Variables

The realistic representation of the steam expansion line in a comprehensive optimiza-

tion study requires including integer variables to define the stage at which each bleed

is being extracted. The steam expansion line constitutes of different turbine stages

with different isentropic and mechanical efciencies, as well as steam leaks. Therefore,

modeling the expansion line requires taking into account for each stage separately,

and integer variables are required in order to specify the extraction stage of each

bleed. A detailed explanation of the procedure and proof of its validity is presented

in [24].

81
2.5.4 Parameters Considered constant

Similar to the arguments presented in [24, 271, for the realistic representation of the

operating units in the power cycle, some variables were excluded from optimization.

The behavior of some components, like the expansion turbines, change in a compli-

cated and component-specific manner when their inputs change. The model here does

not include these dependencies mainly because they are very difficult to estimate. As

a result, parameters like the temperature and pressure of the feedwater exiting the

HRSG/entering the turbine, and more importantly, the reheat extraction pressure

and the reheat delivery temperature are not incorporated as optimization variables.

Moreover, the purity of the oxygen stream resulting from the ASU affects efficiency

and capital cost significantly and thus would be interesting to optimize, but requires

a more elaborate model for each of the ASU and CSU which is beyond the scope of

this work.

2.5.5 Active Constraint Optimization

This study utilizes recent methodological proposals in [24, 28, 29, 27] and in Chap-

ter 1, where it is proven analytically that optimal operating conditions of the cycle

are obtained at some active constraints. More specifically, [24, 27] proves that heat

exchangers need to operate at the MITA specification for optimal performance. A

more dedicated proof for the optimum operation of regenerative Rankine cycles is

presented in [28, 29] and Chapter 1 along with elaborate numerical case studies.

The optimum regeneration necessitates the existence of a double-pinch, i.e. MITA

encountered at the onset of the bleed condensation and simultaneously at the drain

outlet. Therefore, variables can be manipulated at the simulation level to achieve

the desired value of the constraint. The advantages are numerous including reducing

violations and fatal errors in the simulation, constraint violations in the optimization,

and the size of the optimization problem. More importantly, the procedure partially

82
avoids convergence to suboptimal local optima or, even worse, saddle points by guar-

antying that the manipulated variables are set to the values that are obtained at the

global optima. Moreover, the procedure developed is based on explicit equations and

assignments eliminating the need for a spatially distributed model further reducing

computational expense [28, 29].

The variables and constraints coupled in this study are:

1. MITADCSC-HX/DCSC-HX

2. MITAHRSG/rhFW,Main

3. Double-pinchFWH(1 & 2)/QFWH(1 & 2) and rnBLD(1 & 2). The double pinch con-

dition is made up of two simultaneous pinch occurrences requiring the manip-

ulation of two variables. Therefore, both the duty transfer within each closed

FWH and the flowrate of the respective bleeds are defined in terms of the bleed

extraction pressure according to the following equations, [28, 29] and Chapter 1:

r.BLD = M hL(T"'"(PBLD) - AMITAT, PF) - h'(TFi, PEF)


hg,sat(PBLD) - h'(TFi ± AMITAT, PBLD)

Q = rnBLD(hT(PBLD) - h(TFi ± AMITAT, PBLD))

Note that since the expressions are explicit in mBLD and Q there is a unique
double pinch for a given extraction pressure PBLD-

4. qDeaerator/mBLD3. The equality constraint of saturation at the deaerator tank is

satisfied by the deaerator bleed flowrate as proven in [28, 29] and Chapter 1

5. PDeaerator: For optimal operation, the deaerator pressure has to be equal to

the pressure of the deaerator bleed, BLD3, at the deaerator inlet; [28, 29] and

Chapter 1. Therefore, PBLD3 and Pbeaerator are coupled to be equal at the

level of the deaerator, i.e., after accounting for friction and hydrostatic pressure

changes. The deaerator bleed blowrate also plays a role in the optimum value

83
of the deaerator pressure since it affects the amount of pressure loss in the

connection pipes

6. MITADCSC-HX/QDCSC-HX: The allowed minimum internal temperature approach

on the DCSC-HX is achieved by the amount of thermal energy transfer in the

DCSC-HX

7. Balanced DCSC-HX/rhRWSep-in: For optimal operation, the DCSC-HX has to

be thermally balanced and this is satisfied by selecting the recirculating water

flowrate entering the separation column, as proven in Appendix B. There is

no underlying technical or economical limitation for this condition, instead the

constraint is imposed and targeted to ensure the optimal flowrate

2.6 Results

Simultaneous multi-variable optimization of the DCSC flowsheet is performed with

initial conditions set identical to the optimized results of the RHE flowsheet, i.e.

the basecase. Table 2.4 summarizes the optimal RHE, the DCSC basecase, and the

optimal DCSC operation. As expected, changing the thermal recovery unit from a

RHE to a DCSC without optimization reduces the performance substantially, to an

efficiency of 30.9%, a 3.5% points lower than the optimized RHE flowsheet. However,

upon multi-variable optimization the DCSC achieves a surprisingly high performance

of 34.10%, very close to the optimum RHE performance, with considerable difference

in some variables.

To better assess the effect of the operating pressure and importance of thermal

recovery, a pressure parametric optimization is performed and results are plotted in

Figure 2-4. Starting from the optimum found at 12.8bar, operating pressure is in-

crementally varied while optimizing for all variables except pressure. Figure 2-4 also

shows the pressure sensitivity of the DCSC as well as the pressure parametric opti-

84
mization and pressure sensitivity of the RHE, obtained from [24, 27] with an update

to the physical properties of the condensing flue gas. Pressure sensitivity without

optimization does not utilize the complete advantages of the cycle and masks the op-

timal operating conditions. With multi-variable optimization of the DCSC flowsheet

the cycle performance is significantly higher than that of the parametric sensitivity

at any operating pressure. Multi-variable optimization for the DCSC cycle is even

more important than that for the RHE cycle, deduced from the larger improvement

in efficiency of the parametric optimization compared to the parametric sensitivity

in the case of the DCSC cycle. This is because in the DCSC, besides the combus-

tor's operating pressure, the recirculating water flowrate is another very important

optimization variable that play a huge role in achieving the optimal thermal recov-

ery. The other optimization variables also contribute in attaining this high efficiency

performance as explained later.

Compared to the RHE parametric optimization, the DCSC parametric optimiza-

tion is more sensitive to operating pressure near the optimum. The RHE allows for

significant flue gas recovery at low operating pressures, coinciding with the minimum

compression power pressure range. The power requirements in a RHE flowsheet are

insensitive to the operating pressure near the minimum compression requirements

range, [24, 27], resulting in an insensitive range of optimum operating pressure. How-

ever, in the DCSC flowsheet at relatively high operating pressures, where recovery

is significant, pressure losses and compression requirements are very sensitive to the

operating pressure resulting in a more sensitive parametric curve. Prior to the RHE

optimum, the efficiency of the RHE parametric optimization increases rapidly due

to the simultaneous decrease in compression requirements and increase in recovery.

Post the RHE optimum, the efficiency decreases as a result of increase in pressure

losses while recovery plateaus. Prior to the DCSC optimum, the recovery and pres-

sure losses contributions are in opposite directions resulting in the slow increase in

85
efficiency. Post the DCSC optimum, recovery reaches a plateau and the increase in

the pressure losses results in a decrease in the efficiency in almost the same manner

as seen in the RHE at the same operating pressures.

Note that at pressures higher than approximately 12.8bar the performance of

both cycles is very similar. This is because both configurations experience identical

pressure losses and the two different recovery units transfer most of the flue gas's

recoverable latent energy. In fact, at these high pressures the DCSC model marginally

outperforms the RHE model. This is attributed to the temperature of the flue gas

exiting the separator column being lower than that exiting the RHE resulting in

marginally lower compression requirements as discussed in Section 2.6.3.

2.6.1 Variables at Optimal Operation

The high performance of the DCSC flowsheet is achieved by finding the optimum

value of the variables as discussed below.

Combustor Pressure, PComb

The most significant change in the optimal variable values is the combustor operating

pressure. Unlike the RHE, at low operating pressure the amount of water condensed

and thermal energy recovered by the separation column is very small. Therefore,

increasing the operating pressure allows more condensation by increasing the flue

gas dew point leading to higher recovery in the separation column. The optimum

operating pressure for the DCSC flowsheet is found to be 12.8bar, a tradeoff between

recovery and pressure losses.

Recirculating Water Flowrate, rhRW-Sep-in, and Recovered Thermal Energy,QDCSC-HX

The recirculating water flowrate also increases from 100kg/s in the basecase to 190kg/s

in the optimized DCSC flowsheet. This flowrate contributes in effectively condensing

86
the water vapor from the flue gas in the separation column and efficiently transferring

the absorbed thermal energy into the working fluid of the Rankine cycle. The amount

of condensed water vapor in the separation column increases from 10.5kg/s in the

basecase to 33.5kg/s in the optimized results. At optimum operating conditions the

DCSC-HX is balanced i.e., the hot and cold streams have equal thermal capacity rates,

allowing for better thermal matching and larger recovery, as proven in AppendixB.

The recirculating water and the feedwater florwates at the DCSC-HX are 223kg/s.

The former is the sum of the recirculation water original flowrate, 190kg/s, and the

condensed water flowrate, 33.5kg/s. The optimality criteria of a balanced DCSC-HX

while operating at the MITA allow for a high thermal energy recovery by the DCSC

unit, QDCSC-HX = 118MW, close to that recovered in the optimized RHE model, QRHE

= 121MW.

Combustor Duty, QComb

The duty transferred from the combustor to the working fluid of the Rankine cycle,

QComb, again attains the minimum allowed value, 1MW. As detailed in [24, 271, the
combustor losses are of high quality. Minimizing the losses results in higher heat

addition to the high temperature section of the Rankine cycle/ HRSG as opposed to

the low temperature section. Consequently, irreversibilities due to large temperature

differences between the source and the destination are minimized.

Bleeds' Extraction Pressures and Flowrates, PBLD(1,2,3) & rnBLD(1,2,3)

The deaerator operating pressure, PDeaerator, decreases slightly because the thermal

recovery at the low temperature section for the DCSC is a little lower than that of the

RHE. With the slightly cooler feedwater entering the deaerator, the deaerator bleed

flowrate, rnBLD3, increases to ensure saturation at the deaerator tank. Bleed3 extrac-

tion pressure increases slightly due to the larger pressure drop associated with the

87
larger bleed flowrate within identical connection pipes. As expected, the extraction

pressures for the closed FWHs decrease since the temperature of the feedwater exiting

the deaerator is now lower. The bleed flowrates and duty transferred in the closed

FWHs are those that guarantee a double pinch following [28, 29] and Chapter 1.

Main Working Fluid Flowrate, rnFW,Main

The temperature of the feedwater entering the HRSG in the DCSC is slightly lower

than that obtained in the RHE flowsheet due to the changes in the recovery section

(lower bleeds extraction pressures). Lower extraction pressures of bleeds in general

allow lower temperature rise of the feedwater, which is initially colder at the inlet

of the closed FWH compared to the RHE flowsheet. As a result of the cooler FW-

HRSG-in, a unit flowrate of the feedwater entering the HRSG now requires more

thermal energy from the flue gas to elevate its temperature to the high pressure

turbine delivery temperature (600'C). Thus, the amount of working fluid able to

circulate in the power cycle without violating the HRSG pinch specification is lower

for the DCSC than that of the RHE. The optimum value of ThFW,Main drops form 306

kg/s in the RHE optimum operation to 286kg/s in the DCSC optimum operation.

Smaller working fluid flowrate results in lowering the work output.

2.6.2 Flue Gas Pressure Losses

The pressure losses in the recirculation pipes and in the HRSG are higher with the

higher operating pressure of the DCSC flowsheet as seen from the pressure drop

equations, equations (2.1) & (2.2). This translates in higher recirculating fans and

ASU+CSU compression requirements. Thus, the efficiency of the DCSC is somewhat

lower as compared to that of the RHE even at optimum conditions. Note that the

ASU and CSU compression requirements are considered together when comparing the

two flowsheets which operate at different optimum pressures. The ASU requirements

88
in the DCSC flowsheet is larger since it delivers the oxygen at a significantly higher

pressure, but the CSU requirements in the DCSC flowsheet is smaller since it is

compressing the flue gas starting from a significantly higher pressure. However, the

sum of the ASU and the CSU requirements are higher for the DCSC flowsheet due

to the larger pressure losses occurring between the two compression processes (in the

HRSG).

2.6.3 Capital Cost Reduction

Another interesting result is that at optimum operating conditions the DCSC flow-

sheet does not require a chiller. The temperature of the flue gas exiting the separator

column is even lower than that obtained in the RHE model. This result depends

on the specified MITA specification of the RHE (7.1 0 C) and the DCSC-HX (5C),
which is the shortcut method of specifying the acceptable size constraints of heat

exchanges. Heat transfer between two liquids, like the one seen in DCSC-HX, is more

effective than that between a liquid and a gas, like the one in the RHE, and therefore,

a smaller MITA for the former is possible without a larger surface area and addi-

tional capital costs; it follows that a MITA of 5YC for the DCSC-HX is considered

conservative. Note that optimal performance does not require a chiller signifying

lower capital costs of the DCSC, a result that is not likely to be discovered without

a comprehensive optimization.

The DCSC behavior suggests more chances for capital cost reduction. The op-

erating pressure of the cycle at optimum operating conditions is close to the design

pressure of the NOx and SOx purification columns in the CSU. This provides two

additional benefits. First, the first CSU compressor and its intercooling can be elim-

inated since the flue gas pressure is already suitable for the purification processes.

Second, process intensification is possible (at least in principle), whereby purifica-

tion can occur simultaneously with the recovery process in the separator column of

89
the DCSC eliminating the need for additional purification columns and reducing the

capital cost and size of the CSU and the power plant. The increase in practicality

and reduction in the capital cost associated with the DCSC model increases the com-

petitiveness and interest in this pressurized OCC process, and is a topic of future

work.

2.6.4 Validation of the Optimization Results

The validation of the results is performed by perturbing the values of the optimization

variables and examining the objective function. Notice that numerical errors are only

relevant to variables that are not governed by criteria of optimal operation. Mean-

while, the variables that are defined by the active constraints and optimal criteria are

automatically set to their exact optimal value. Error analysis shows that the pertur-

bation of the variables does not increase the value of the objective function and that

the results are indeed optimal. Moreover, results also show that the objective func-

tion is insensitive to the extraction pressures close to optimum, a similar behavior is

observed in [28, 291 and in Chapter 1. For example, even a 10% change in the value of

the bleeds' extraction pressures results in less 0.03 percentage points decrease in the

overall efficiency; the insensitivity to the extraction pressures motivated a detailed fu-

ture work study of the pressurized OCC process's flexibility, [35] and Chapters 3&4,

and suggests that retrofitting may be viable. The sensitivity of the objective to the

other variables is also studied. Around the optimum solution, the variables that are

most influential are the combustion pressure, PComb, the bleeds flowrates, mhBLD, the

feedwater heaters and DCSC-HX duty transfer, QFWH & QDCSC-HX, and the recycling

water flowrate. Note that among the sensitive variables, only the combustion pressure

is not governed by an optimal criterion of operation.

90
2.7 Model-Based Optimization and Effect of De-

sign Assumptions

It is very important to state that the design considerations and assumptions, i.e.,

the parameters that are considered fixed in the process, are crucial in determining

the behavior and the response of the cycle. In this study a fixed design and a set

of economical considerations for the HRSG and the recycling pipes are enforced,

Section 4.4.2, and the pressure losses are calculated by similarity analysis. This

causes an increase in the pressure losses at an accelerating rate with the increase

the operating pressure above 6.25 bars, [27, 24]. Therefore, the increased flue gas

recovery at higher pressures is faced with increased pressure losses, and after reaching

almost full thermal recovery the efficiency of the process starts to decrease rapidly,

as seen in the parametric optimization curve of Figure 2-4.

On the other hand, changing the design considerations results in altering the

response of the process with increasing the pressure. For example, in [68] a similar

flowsheet is studied but with different design parameters. Therein, the pressure losses

for the HRSG and the recycling pipes are assumed to be a constant fraction of the

operating pressure and in contrast to this chapter do not increase at an accelerated

rate. Therefore, therein, the parametric optimization curve exhibits a different behav-

ior, presented in Figure 2-5; first, below 20 bar the efficiency of the process increases

rapidly with an increase in the operating pressure due to the increased thermal recov-

ery. Second, in the range of 20 to 30 bar, the efficiency is almost constant. Finally,

above 30 bar, the efficiency decreases at a very slow rate with increasing pressure.

In addition, in [68] two other recycling configurations are studied. The first is

dry-recycling, where flue gas is recycled after thermal recovery and water condensa-

tion. If safety measures against particulate agglomeration in the recycling pipes are

not considered, then dry-recycling requires significantly lower recycling flowrates and

91
power consumption by the recycling-fans because the recycled flue gas is cooler, has

a smaller flowrate, and contains almost no water vapor. However, the flue gas from

dry-recycling is heated by at least 60'C as a safety measure against particulate ag-

glomeration. This reduces the savings in the compression power of the recycling-fans

by increasing the temperature, the flowrate, and the pressure losses of the recycled

flue gas in the recycling pipes. The heating of the dry-recycled flue gas is performed

by the flue gas exiting the HRSG/entering the DCSC. Moreover, compared to the

wet-recycling, results show that in dry-recycling a larger fraction of flue gas enters

the DCSC and transfers a larger fraction thermal energy at the low-temperature sec-

tion on the account of the thermal energy transfer at the high-temperature section;

the exergy of the heat transfer process to the working fluid in dry-recycling is lower

than that of the wet-recycling. As a result, dry-recycling configuration has lower

efficiency than that of wet-recycling; also dry-recycling requires a larger separation

column increasing to the capital cost. Note that the Isotherm combustor requires high

rates of flue gas recycling for the mild combustion process, [30], which may not be

satisfied by dry-recycling, and therefore a dual-recycling configuration is presented as

a compromise between the wet and dry-recycling configurations. The second studied

configuration is appropriately named by the authors dual-recycling, where the recy-

cled flue gas to the combustor is in the wet-recycling form, while the flue gas recycled

to the HRSG is in the dry-recycling form. The exergy of the heat transfer to the

working fluid, and the power requirements of the recycling fans for the dual-recycling

is between those of the wet-recycling and dry-recycling configurations. However, for

the assumption of fixed fraction of pressure losses in [68] the power savings on recy-

cling are not sufficient to compensate for the exergy reduction of the heat transfer;

therefore, the efficiency of the dual-recycling is less than that of the wet-recycling.

Although not evaluated, dual-recycling is more promising when considering the design

criteria herein because pressure losses are larger and more sensitive to the operating

92
pressure. This stresses again on the importance of the design criteria and assumptions

considered for the process.

Comparing the different behavior and response of the process for the different

design criteria helps realize a very important fact: Model-based optimization is of

utter most importance. Optimization is not only needed to discover the optimal per-

formance of a given process, but also to assess the influence of the different design

assumptions and considerations. Without optimization, different design criteria may

misleadingly result in similar performance but masking the behavior, response, and

sensitivity of a certain process. Optimization allows discovering (i) the most favor-

able design consideration for a given application, (ii) the optimal operation for the

considered design conditions. and (iii) the overall behavior, response, and sensitivity

of the process at any given operation particularly the optimal operation. Ideally,

deciding on the set of design criteria to implement should come after the conclusions

derived from the optimization of all the candidate sets of design criteria. While this

seems computationally very demanding, it comes with a relatively small additional

effort. The model setup of the process and the optimization formulation are almost

identical between the different sets of the considered design criteria.

2.8 Conclusion

Pressurized OCC can mitigate the problem of emissions accompanying the combustion

of an abundant and cheap fuel. Moreover, the comprehensive optimization applied

to the process results in performance and efficiency much higher than those of simu-

lations and single variable manipulations and at operating pressures more attractive

than that proposed in literature. With the help of recently proposed active con-

straint optimization concepts, this study evaluates the performance using a separator

column, a more practical heat recovery unit than a surface heat exchanger/RHE.

93
Upon comprehensive optimization, a pressurized OCC process utilizing a direct con-
tact separation column/DCSC does not suffer large efficiency decrease compared to

an optimized pressurized OCC utilizing an RHE. Optimization results in attractive

performance of the OCC process utilizing a DCSC along with lower capital cost than

what simulations or sensitivity analysis result in. The combustor operating pressure

increases from the optimum range of 7.41 bar for the RHE flowsheet to 12.8bar for the

DCSC flowsheet. The larger operating pressure mainly increases the cost of the com-

bustor and the recycling pipes, however, capital costs may be offset by the major cost

reductions in the flue gas drying and cleaning processes by eliminating a no-longer-

needed compressor with intercooling and by process intensification. All optimization

variables play an important role in achieving the attractive process performance when

simultaneous multi-variable optimization is performed.

Moreover, the study results in an important conclusion that model-based op-

timization is essential. Different design assumptions lead to significantly different

response and behavior, and optimization is not only needed to discover the optimal

performance of a given process, but also to assess the influence of the different design

assumptions and considerations. The important lesson learned is that deciding on the

set of design criteria to implement should come after the conclusions derived from

the optimization of all the candidate sets of design criteria.

94
Pressurized - ASU
02 02
ompressorH

N2

-r c-pri FW-HRSG-in
FG- ec-seEr- - -

om-Gas-in k with (x fo th Recf RHE or DCSC r su


at c t Recovery Unit -e P.kc Sequestrated C02
Coa Cfmbustio

Slurry Gaoo- G as G-Recov-out -

Te ure FG-Recov-in 16f1 -7_+

ad vented gas
Ash Controller
FW-Recov-in
AtomizerCondensate

i HT .- - I - 1
RZ leed1 Bleed

Rehjeat -- IP 9zmr Cooling Water

FWHRG-n FW FW Deaerator Condenser

HP-Pump

LP-Pump

Figure 2-3: 17 Optimization variables in purple marked with (o), and 14 constraints
in green marked with (x) for the DCSC configuration. The RHE configuration is used
for later chapters.
X: 7.41
Y: 34.41
34.5 r X: 12.8
Y: 34.1

34 F

33.5

0 33 Direct contact separation column (DCSC)


-J
parametric optimization
32.5 - Surface heat exchanger (RHE)
parametric optimization
Direct contact separation column
32 sensitivity analysis
u.w
Surface heat exchanger
31.5 sensitivity analysis
I M i MI M j M I ft I M I I

31

30.5
-

0
/ \#

5
~

10 15 20 25
Combustor Pressure (bar)

Figure 2-4: RHE and DCSC pressure parametric optimization and pressure paramet-
ric sensitivity result
Table 2.4: Key results of the base-case, the pressure only parametric study, and opti-
mization runs

Variable Optimization RHE DCSC with RHE variables Optimization


DCSC (no
chiller needed)
Fuel Flowrate 30.00kg/s 30.00kg/s 30.00kg/s
Slurry water 16.50kg/s 16.50kg/s 16.50kg/s
flowrate
Atomizer 2.50kg/s 2.50kg/s 2.50kg/s
Stream flowrate
Air flowrate 311.2kg/s 311.2kg/s 311.2kg/s
PComb 7.41bar 7.41bar 12.8bar
Objective Func- 34.4% 30.9%" 34.1%
tion (LHV)
QComb 1MW 1MW 1MW
mFW,Main 306kg/s 277kg/s 286kg/s
PBLD1 99.Obar 99.0bar 65.7bar
rnBLD1 62.2kg/a 62.2kg/s 36.8kg/s
PBLD2 26.Obar 26.Obar 25.2bar
rnBLD2 14.7kg/s 14.7kg/s 17.0kg/s
PBLD3 16.8bar 16.8bar 14.7bar
rnBLD3 9.27kg/s 9.27kg/s 9.52kg/s
PDeaerator 14.7bar 14.7bar 12.2bar
QFWH1 138MW 138MW 78.4MW
QFWHM2 37.7MW 37.7MW 44.1MW
TFG-RHEout 36.9 0 C N/A N/A
TFG-Sep-out N/A 123.7 0 C 35.8 0 C
RW-Sep-in N/A 100kg/sb 190kg/s
QDCSC-HX N/A 44.9MW 118MW
QRHE 122MW N/A N/A
Dependent Variables
TCool-Gas 327 C0 289 0 C 292 0C
TFW-HRSG-in
0
322 C 283 0 C 2870C
Tcomb-Gas-in 309 0C 279 0 C 290 0 C
Flue gas flowrate 1,138kg/s 1,069kg/s 1,081kg/s
in HRSG
Condensed Wa- 33.2kg/s N/A N/A
ter RHE
Condensed Wa- N/A 10.5kg/s 33.5kg/s
ter Sep-Col
APHRSC 0.265bar 0.275bar 0.8654bar
APpip, primary 0.054bar 0.046bar 0.088bar
AP, sec- 0.035bar 0.039bar 0.088bar
ondary
Prim. recycling 281kg/s 274kg/s 276kg/s
pipe flowrate
Sec. recycling 737kg/s 675kg/s 684kg/s
pipe flowrate
a A Chiller is required for this choice of variables. If no chiller was incorporated the objective
function would be even lower, 30.27%
b If ~ i, ~ +' 1'9llr /~pcc rn+___indf-~ mrrdncf 1~TC~
33.0
11
jj111111111111111111

32.0

C
0 31.0

30.0

29.0
0 10 20 30 40 50 60
Process operating pressure (bar)

Figure 2-5: Parametric optimization results for the pressurized OCC utilizing a DCSC
with different set of design assumptions and considerations, namely the pressure losses
are a constant fraction of the operating pressure, [68]
Chapter 3

Pressurized Oxy-Coal Combustion:

Ideally Flexible to Uncertainties

3.1 Summary

Simultaneous multi-variable gradient-based optimization with multi-start is performed

on a 300 MWe wet-recycling pressurized oxy-coal combustion process with carbon

capture and sequestration, subject to uncertainty in fuel, ambient conditions, and

other input specifications. Two forms of flue gas thermal recovery are studied, a sur-

face heat exchanger and a direct contact separation column. Optimization enables

ideal flexibility in the processes: when changing the coal utilized, the performance

is not compromised compared to the optimum performance of a process specifically

designed for that coal. Similarly, the processes are immune to other uncertainties like

ambient conditions, air flow, slurry water flow, atomizer stream flow and the oxidizer

stream oxygen purity. Consequently, stochastic programming is shown to be unneces-

sary. Close to optimum design, the processes are also shown to be insensitive towards

design variables such as the areas of the feedwater heaters. Recently proposed ther-

modynamic criteria are used as embedded design specifications in the optimization

99
process, rendering it faster and more robust.

3.2 Motivation

Besides performance and capital cost, another characteristic that favors adopting a

power generation process is its flexibility regarding changes in inputs, operation pa-

rameters, and desired output. Coal type and specifications vary significantly from one

source to another, or even from different batches of the same source. A process that is

optimized for a given coal type would be unattractive if its performance deteriorates

or suffers when that specific coal is not economically attainable. With the change

of coal type, other input specifications and parameters need to change accordingly,

like the oxidizer flowrate, the flowrate of slurry water responsible for transferring the

coal into the combustor, and the atomizer stream flowrate, which contribute to the

alteration of the cycle's behavior. Other forms of uncertainty are due to the am-

bient conditions, in particular that the cooling temperature may vary significantly.

Additionally, it is desirable to have a process that is insensitive in the selection of

feedwater heaters (FWHs) areas, which is of high interest for retrofitting. Finally,

change of load is also a significant form of uncertainty and is discussed in Chapter 4.

The aim of this chapter is to find a flexible design for the pressurized OCC pro-

cesses presented in [24, 27, 38] and in Chapter 2 with respect to changes in coal type,

ambient conditions, input streams flowrates, and oxidizer stream's oxygen purity. The

models involve rigorous accounting of irreversibilities and losses in order to accurately

assess the performance compared to other CCS technologies and in order to accurately

evaluate the pressurized OCC flexibility. Section 3.3 summarizes the flowsheets and

the models of two aforementioned pressurized OCC processes. Section 3.4 discusses

the implemented optimization formulation: the objective, variables, and constraints,

in particular the active constraint optimization process based on thermodynamic cri-

100
teria. Section 3.5 evaluates the processes' flexibility, demonstrating ideal flexibility

regarding coal and FWHs areas.

3.3 Flowsheet and Model Description

The RHE and the DCSC processes and models studied here are identical to the ones

presented [24, 27, 38] and in Chapter 2, and should be referred to for further details.

AspenPlus@ is used for modeling the flowsheets.

Figure 2-3 shows a schematic of the processes and includes tags for the variables

and constraints required for the following sections.

The RHE is a counter-current surface heat exchanger where the hot stream is the

flue gas and the cold stream is the working fluid of the Rankine cycle. In contrast

the DCSC, seen in Figure 2-2, utilizes an intermediate recirculating water stream

between the flue gas and the working fluid of the Rankine cycle. The recirculating

water transfers the thermal energy from the flue gas to the Rankine cycle in a liquid-

liquid heat exchanger (DCSC-HX). The figure also includes tags for the variables

and constraints required for following sections. Refer to Section 2.3 for a detailed

explanation of the DCSC recovery unit.

The coal variations are represented by using two substantially different coals shown

in Table 3.1. CoalA, used in [19, 20, 21, 23, 24, 27, 38], is a typical bituminous coal

with composition similar to Venezuelan and Indonesian coals. CoalB, a south African

coal almost identical to Douglas Premium or Kleincopje coal, is of a substantially

lower quality. The flowrates of coal are set to keep the heating rate (the product of the

the flowrate and the lower heating value) constant. The properties and constituents

of each coal are important since they specify the amount of coal needed for a given

heating load, the amount of oxidizer needed, the flowrates of slurry water and atomizer

streams, and the composition of the flue gas.

101
Table 3.1: Mass dry basis specifications of high, (CoalA), and low, (CoalB), quality
coals. Moisture is based on moisture-included basis.

Coal Type CoalA CoalB


Coal mass flowrate 30.00kg/s 35.53kg/s
Coal HHV 31.09MJ/kg 26.42MJ/kg
Coal LHV 29.88MJ/kg 25.23MJ/kg
Moisture 6.4 7.4
Ash 7.479 14.7
Carbon 75.962 71.2
Hydrogen 5.021 3.9
Nitrogen 1.282 1.7
Sulfur 0.534 0.6
Oxygen 9.722 7.9

3.4 Optimization Formulation

3.4.1 Optimization Objective

As aforementioned, several sources of uncertainty are considered. For simplicity,

the methodology followed is first described for the coal input uncertainty and then

repeated for the other sources of uncertainties.

To find the optimal design and operation, the uncertainty of the coal input must

be accounted for. The uncertainty influences, at least in principle, both the optimal

design and the optimal operation. A common design has to be used for all coals, while

operation can be adjusted from coal to coal. One method to address this is stochastic

optimization, [41, 42]. The objective is to obtain a flexible cycle that maximizes

the expected value of efficiency for an expected distribution of the uncertain input

specifications and parameters. The optimal process depends, at least in principle,

on the distribution of uncertain parameters, and therefore care has to be applied in

identifying this distribution.

In the ideal case, the optimal performance of the flexible process for each coal is

as good as the performance of the best process for this coal. In that case, not only

102
maximal performance is achieved, but also there is no need to estimate an expected

distribution of the uncertain specifications. The values of the design variables in

that ideal scenario coincide with solutions for the stochastic optimization problem.

Moreover, they are also solutions for the hierarchical optimization problem, presented

in [43, 44, 451, where the objective functions are the performance of each coal.

As will be explained in further details later, herein, first, the processes are designed

for each of the two coals, then the coal is changed. Subsequently, optimization,

using recently proposed optimality criteria, of the operation is performed on the

fixed design; the obtained performance is compared to that of the process designed

specifically to the implemented coal. As will be demonstrated, some designs that do

not take flexibility into account can results in significantly suboptimal performance

with the change in the input specifications, while other designs satisfy the desired

ideal scenario within a maximum discrepancy of 0.02%. Although this discrepancy

is already very small, hierarchical optimization performed in Section 3.5.2 eliminates

the discrepancy and also uncovers flexibility regarding parameter specifications like

FWHs areas; ideal flexibility regarding the other uncertainties are also presented and

the results are presented and discussed.

3.4.2 Design and Operation Variables

The decision variables are characterized as design or operation. Each design variable

acquires a unique value invariant among the different plant operations. In contrast,

an operation variable can change, within a certain range, between the different plant's

operations. Table 3.2 summarizes the variables while Figure 2-3 marks the variables

on the flowsheet. The methodology of implementing active constrains depends on the

variables' characterizations and is discussed in Section 3.4.4.

The combustor's operating pressure, PComb, ., is an operation variable while the

maximum allowed pressure or the design pressure, PComb, d is a design variable.

103
Herein, an upper bound of 30bar is considered on the design pressure but substan-

tially lower optimal pressures are expected and observed. With a lower quality coal,

the water content in the flue gas increases, and therefore the pressure is favored to

increase in order to enhance the thermal recovery. The design pressure is the maximal

pressure among all operating pressures.

The combustor's duty, QComb, is a design variable since it relates to the design

of the combustor and the refractory insulation installed. Note that the combustor

temperature is fixed.

In principle, the main feedwater flowrate can be changed anytime during the oper-

ation, however, the efficiency of the turbines deteriorates, usually very substantially,

if the flowrate is varied significantly from the design flowrate. Therefore, rFWMain

is considered to be an operation variable under the condition that the optimal value

does not vary a lot between the different operations. The optimal results show that

the flowrate change is acceptably small, within few percent demonstrating that there

is no need for performance curves of the turbine.

The bleeds' flowrates can change with different coals and thus considered as op-

eration variables. However, the bleeds' extraction pressures are fixed with the ini-

tial design of the turbine expansion line and thus are design variables. Similar to

[24, 27, 38] and Chapter 2, integer variables are used to select the extraction stage,

extraction positions; these are also design variables.

The duty transferwithin each FWH is an operationvariable. It varies with changes

in other operation variables like bleed and feed flowrates, as well as changes in design

variables like extraction pressures. Constraint specifications and their values also play

a role in the value assigned to the FWHs duty, like the minimum internal temperature

approach (MITA) and/or area specifications.

The deaerator operating pressure is an operation variable determined primarily

by the low pressure pump delivery pressure and the deaerator bleed pressure at des-

104
tination. The pump and deaerator pressures are allowed to vary but also within

relatively small ranges. The maximum allowable deaerator pressure is a design vari-

able dictated by material and structural properties, but the optimum is safely below

the maximum allowed pressure. The deaerator pressure is not expected to vary signif-

icantly since the deaerator bleed pressure is a design variable, and since the deaerator

bleed flowrate is an operation variable. In optimization, an excessively large range

of the deaerator pressure is implemented [0.1-30] in order to investigate any perfor-

mance improvements. However, results show only minor variations in the optimum

deaerator pressure with the different operating conditions and all are well within the

practical ranges of [1.5-20]bar.

Additional Variable Specific for the RHE Flowsheet

The temperature of the flue gas exiting the RHE, TFG-RHE-out, is also an operation

variable dictating the amount of recovered thermal energy. However, it is not expected

to vary between the two coals especially since the exchanger operates under a fixed

MITA specification of 7.5'C. For simplicity the RHE is modeled assuming a constant

MITA instead of a constant area. This is acceptable bevause the variations in the

streams of the RHE are relatively minor at optimal conditions. Moreover, the MITA

is relatively large resulting in small variations in the required area with the minor

changes in the streams conditions.

Additional Variables Specific For the DCSC Flowsheet

The duty transfer in the DCSC-HX, QDCSC-HX, is an operation variable. Similar

to the RHE the MITA is used for the specification on the DCSC-HX. Moreover,

similar to the main feedwater flowrate, the recirculating water flowrate, TRW-Sep-in, is

an operation variable as long as its value does not change significantly between the

different operations; the flowrate affects the sizing of the heat exchanger, separation

105
column, pump, connection pipes etc. The values of QDCSC-HX and rnRW-Sep-in axe not

expected to change between the different operations of the DCSC flowsheet especially

since both variables are shown to have their optimum value at active constraints (see

Section 3.4.4); the results validate this assumption.

To accurately model the operating units in the power cycle, some variables like

the oxygen purity, feedwater temperature and pressure at the exit of the HRSG, and

Reheat temperature and pressure, are excluded from optimization, similar to [24, 38].

3.4.3 Constraints

Constraints on the admissible design and operation are imposed based on physical,

practical, and economical considerations. Each process has ten constraints, nine of

which are common. The constraints are listed in Table 3.3 and illustrated in Figure 2-

3.

3.4.4 Active Constraint Optimization

In [24, 27, 28, 29] and Chapters 1&2 it is proven that optimal operating conditions

occur at some active constraints. Enforcing these constraints as operation specifica-

tions facilitates the optimization in several aspects: (i) avoid constraints violations,

(ii) avoid simulation errors and failures, (iii) accelerate convergence, (iv) avoid con-

vergence to suboptimal local optima. The desired active constraints can be satisfied

at the simulation level by manipulating the main influencing variables.

The following constraints and variables are coupled:

1. MITAHRSG/rhFW, Main: The allowed minimum internal temperature approach

constraint on the HRSG is achieved by manipulating the main feedwater flowrate

2. & 3. Double-pinchFWH(1&2)/QFWH(1&2) and rnBLD(1&2). Both the duty trans-

106
Table 3.2: Design and Operation Variables. The deaerator operating pres-
sure should be above atmospheric, but here the lower bound was intention-
ally taken sub-atmospheric to examine if it would lead to any advantage
in performance. The variables, ThFW, Main & PDeaerator & ?nRW-Sep-in, are
considered operation variables but expected to vary a little between the
different coals; results satisfy this condition

Number Variable Range Variable Type

1 PComb, d [1.283 - 30] bar Design


2 PComb, o [1.283 - PComb, d] bar Operation
3 QComb [1 - 30] MW Design
4 TnFW, Main [240-340] kg/s Operation
5 BLD1_stage integer Stages: 1-4 Design
variable
6 PBLD1 [250 - 30] bar Design
7 ThBLD1 [0 - 60] kg/s Operation
8 BLD2-stage integer Stages: 3-6 Design
variable
9 PBLD2 [120 - 10] bar Design
10 ThBLD2 [0 - 30] kg/s Operation
11 BLD3_stage integer Stages: 5-7 Design
variable
12 PBLD3 [30 - 4.5] bar Design
13 rnBLD3 [0 - 30] kg/s Operation
14 QFWH1 [0 - 200] MW Operation
15 QFWH2 [0 - 200] MW Operation
16 PDeaerator [0.1 - 30] bar Operation
Variables specific for the RHE
17 TFG-RHEout [30 - 150] 0C Operation
Variables specific for the DCSC
18 QDCSC-HX [50 - 150] MW Operation
19 rnRW-Sep-in [50 - 400] kg/s Operation
Table 3.3: Optimization Constraints, [24, 27]. For the initial design
purposes MITAs are taken as the FWHs' constraint. However, for flex-
ibility assessment, fixed sizes are imposed on the FWHs, Section 3.5;
for generality two different sets of FWHs areas, obtained from the in-
dependent design of each coal with MITA specifications on the FWHs,
are assessed for each process.

Number Constraint Value


1 MITAHRSG 3.7'C economizer section
2 MITAFwH1 2.1 0 C
2
RHE-AreaA=7,069m
AreaFWH1 RHE-AreaB=6,728m 2
DCSC-AreaA=6,158m 2
DCSC-AreaB=5,552m 2
3 MITAFWH 2 2.1 0 C
RHE-AreaA=5,254m 2
AreaFWH2 RHE-AreaB=4,525m 2
DCSC-AreaA=4,227m 2
DCSC-AreaB=3,561m 2
4 qDeaerator Saturated Liquid
5 TCool-Gas 20'C above acid condensa-
tion temperature
6 TFW-HRSG-in 5C above acid condensa-
tion temperature
7 TCom-Gas-in 20'C above acid condensa-
tion temperature
8 C02-Cap 94% of total CO 2 produced
required to be captured
9 CO 2 -Pur captured CO 2 is 96.5% pure
Constraint specific to the RHE
10 MITARHE 7.5 0 C
Constraint specific to the DCSC
10 MITADrCSCHX 50 C
10 MIT~nf Qt" -T5C
fer within each closed FWH and the flowrate of the respective bleeds are utilized

and defined in terms of the bleed extraction pressure according to the following

equations, [28, 29] and Chapter 1:

hl(Tsat(PBLD) - AMITA,FWHT, PEW) - h'(TFW,i, PEW)


mBLD = mW hg,sat(pBLD) - h'(TFWi ± AMITA,FWHT, PBLD)

QFWH = rBLD(hT(PBLD) - h(TFwi + AMITA,FWHT, PBLD))

where h is the specific enthalpy, h' is the specific enthalpy of the liquid, Tsat is

the saturation temperature, and hT is the specific enthalpy at a certain point

along the steam expansion line.

4. PDeaerator: For optimal operation, the deaerator pressure has to be equal to

the pressure of the deaerator bleed, BLD3, at the deaerator inlet; [28, 29] and

Chapter 1. Therefore, PBLD3 and PDeaerator are coupled to be equal at the

level of the deaerator, i.e., after accounting for friction and hydrostatic pressure

changes. The deaerator bleed blowrate also plays a role in the optimum value

of the deaerator pressure since it affects the amount of pressure loss in the

connection pipes

5. qDeaerator/rBLD3. The mixture inside the tank deaerator has to reach the satu-

rated liquid state for the effective removal of dissolved air in the working fluid.

This constraint is satisfied by the low pressure bleed to the deaerator

Additionally for the RHE

7. MITARHE/TFG-RH-out: The allowed minimum internal temperature approach

on the RHE achieved by manipulating the temperature of the flue gas exiting

the RHE

Additionally for the DCSC

109
7. MITADCSC-HX/QDCSC-HX: The allowed minimum internal temperature approach

on the DCSC-HX is achieved by the amount of thermal energy transfer in the

DCSC-HX

8. Balanced DCSC-HX/RW-Sep,-in: For optimal operation, the DCSC-HX has to

be balanced and this is satisfied by the recirculating water flowrate entering the

separation column, [38] and Chapter 2.

3.5 Ideally Flexible Process to Coal, FWHs Areas,

Input Flows and Specifications, and Ambient

Temperature

3.5.1 Methodology for Flexibility Assessment

In this section, the method for coal variations flexibility assessment is presented but

the same applies to all the addressed uncertainties. The performance of the process

differs when operating with the different coals due to the different specifications,

particularly in the heating value and the water content. The total heat rate input

of the fuel based on the LHV is held constant for reasons explained later. More

specifically, when operating with CoalB the coal flowrate is set to 35.53kg/s to obtain

a heat rate input into the cycle equal to that provided by a 30.00kg/s of CoalA. A

direct consequence of the coal flowrate are adjustments of several other flowrates to

maintain the proper operation of the combustor and the coal-water slurry delivery.

In Section 3.5.2 results show that this approach leads to maintaining the mFW, Main

close to nominal, as intended, validating that the behavior of the turbine expansion

line, with fixed specifications, is accurately represented.

Deterministic optimization is performed in three steps to assess flexibility; then,

110
hierarchical optimization is performed. It is demonstrated that stochastic optimiza-

tion is not needed. The performed series of runs are summarized in Table 3.4 and

illustrated graphically in Figure 3-1. First in Stepi, optimization is performed with

a MITA specification of 2.1 0 C on the FWHs for each of the two coals; the runs for

CoalA are essentially those presented in [24, 27, 38] and Chapter 2. Second in Step2,

the process is optimized for each coal with the areas of the FWHs fixed to the opti-

mal values from Stepi for the other coal; Step2 determines the optimal design and

performance of the process with the given coal and area specification. Third in Step3,

the processes and designs of Stepi are optimized using only the operation variables

while holding the design variables and FWHs areas fixed to the optimum value of the

other coal. Finally, comparing the efficiency of each coal in Step2 and in Step3, the

flexibility of the process is evaluated.

In other words, the comparison shows whether using a coal in a cycle designed for

a different coal can reach the same performance as one designed specifically for the

former coal. Note that the comparison is performed for two different FWHs' area to

avoid favoring one of the two coals as well as showing that the flexibility results are

not specific to a given FWH size.

Results of Flexibility Assessment

Tables 3.5 through 3.8 show the results of the runs performed for cases RHE-A, RHE-

B, DCSC-A, DCSC-B respectively as defined in Table 3.4. Upon changing the coal

type in a cycle designed for a given coal without changing the operation, the perfor-

mance is significantly lower than the highest possible performance for the given coal.

These results are omitted for brevity. Considering RHE-A-3 in Table 3.6, by opti-

mizing the operation variables while holding the design variables identical to those

of CoalB's optimum performance, the cycle efficiency is essentially the optimum per-

formance attainable by CoalA. However, not all optimum designs of one coal can

111
Table 3.4: The runs performed to check the flexibility of the OCC cycle, with an RHE
or DCSC thermal recovery unit, under fuel uncertainty

Step number Case Identification Case Description


Stepi RHE-A-1 RHE Flowsheet, CoalA, Optimization of design
and operation variables, FWHs' MITA constraint
of 2.1 C
RHE-B-1 RHE Flowsheet, CoalB, Optimization of design
and operation variables, FWHs' MITA constraint
of 2.1 C
DCSC-A-1 DCSC Flowsheet, CoalA, Optimization of design
and operation variables, FWHs' MITA constraint
of 2.1 C
DCSC-B-1 DCSC Flowsheet, CoalB, Optimization of design
and operation variables, FWHs' MITA constraint
of 2.1'C
Step2 RHE-A-2 RHE Flowsheet, CoalA, Optimization of design
and operation variables, FWHs' areas from case
RHE-B-1
RHE-B-2 RHE Flowsheet, CoalB, Optimization of design
and operation variables, FWHs' areas from case
RHE-A-1
DCSC-A-2 DCSC Flowsheet, CoalA, Optimization of design
and operation variables, FWHs' areas from case
DCSC-B-1
DCSC-B-2 DCSC Flowsheet, CoalB, Optimization of design
and operation variables, FWHs' areas from case
DCSC-A-1
Step3 RHE-A-3 RHE Flowsheet, CoalA, Optimization of operation
variables, Design Variables and FWHs' areas from
case RHE-B-1
RHE-B-3 RHE Flowsheet, CoalB, Optimization of operation
variables, Design Variables and FWHs' areas from
case RHE-A-1
DCSC-A-3 DCSC Flowsheet, CoalA, Optimization of opera-
tion variables, Design Variables and FWHs' areas
from case DCSC-B-1
DCSC-B-3 DCSC Flowsheet, CoalB, Optimization of opera-
tion variables, Design Variables and FWHs' areas
from case DCSC-A-1
Stepl
Power Plant designed for Coal A
Coal A
max {efficiency I inputs = CoalA specs}
Get: opt designA, opt operationA, AreaA

'2i

00 (

Step3
Step2 Power Plant designed for Coal A (area
Power Plant wit IAreasA, Optimized specification fixed to AreaA), operation
Coal B (all variables) for coalB Coal B optimized for CoalB
V Max {efficiencyIinputs =- CoaIB
areaA} Compare and Max{efficiency I inputs CoalBopt designA,
}evaluate flexibility areaA}

Figure 3-1: Evaluations performed for flexibility assessment of the RHE process while
using areas favoring the design for CoalA. The same 3-steps procedure is repeated for
areas favoring the design for CoalB and for the DCSC process; (four times in total).
attain the optimum operation when changing the coal as seen in Table 3.5. These

tables signify that the RHE cycle can be ideally flexible. Similarly, Tables 3.7 and

3.8 show that the DCSC process can also also be ideally flexible (discrepancies within

0.02%). For the flexibility evaluations of both processes and for the hierarchic opti-

mization presented next, with current optimization solvers, it is necessary to utilize

the optimality criteria (proposed in [28, 29, 38] and Chapters 1&2) to obtain this

ideal flexibility.

3.5.2 Hierarchical Optimization

The above evaluations show that the cycle is nearly ideally flexible. Hierarchical

optimization is performed to find a set of values for the design variables that mini-

mizes/eliminates the discrepancies between the multi-coal design and the two single

coal designs. Only ideally flexible designs, within the specified allowable tolerances,

are feasible in the hierarchy optimization formulation. This is achieved by introducing

a constraint on the objective function in the formulation below. rq is the efficiency, A

and B stand for the coal type, g is the set of constraints. e is the allowed discrepancy

from the optimum. x is the set of design variables that have to be common among

the different operations of the flexible process, and y is the set of operation variables

that can change between the different operations The results from Section 3.5.1 are

feasible for e > 0.02%.

Table 3.9 shows results of the hierarchical optimization. The procedure is re-

peated for the two processes and for the two respective sets of FWHs area. The

performance of all four optimization cases shown match the maximum performance

of the respective identical cycles deterministically optimized for a single coal. Note

that Column2's performance is higher than that of RHE-A-1 because the former is

optimized with fixed areas specification while the latter with fixed MITAs. The same
Table 3.5: Results for RHE flowsheet fuel flexibility. Area 'favoring' the design of
CoalA; AFWH1=7,069m 2 AFWH2=5,254m 2 . The changes in the values of mFW Main
and PDeaerator are acceptably small, Section 3.4.2.

Variable RHE-A-1 RHE-B-2 RHE-B-3


Input Parameters
Fuel Flowrate 30kg/s 35.53kg/s 35.53kg/s
Slurry water flowrate 16.50kg/s 19.22kg/s 19.22kg/s
Atomizer Stream 2.50kg/s 2.96kg/s 2.96kg/s
flowrate
Air flowrate 311.2kg/s 332.5kg/s 332.5kg/s
Efficiency (LHV) 34.41% 33.06% 32.82%
Independent and Key Dependent Variables
PComb,d 7.41bar 9.67bar 7.41bar
PComb,o 7.41bar 9.67bar 7.41bar
QComb 1MW 1MW 1MW
mFW Main 291kg/s 286kg/s 285kg/s
PBLD1 80.0bar 77.8bar 80.Obar
rnBLD1 45.4kg/s 44.2kg/s 44.3kg/s
PBLD2 26.lbar 26.2bar 26.lbar
rnBLD2 23.9kg/s 19.9kg/s 24.2kg/s
PBLD3 9.44bar 11.4bar 9.44bar
rnBLD3 1.98kg/s 2.12kg/s 0kg/s
PDeaerator 9.30bar 11.4bar 9.02bar
Q0textFWH1 98.2MW 95.4MW 95.7MW
Q0textFWH2 63.6MW 52.1MW 64.6MW
TFG-RHEout 36.9 0 C 36-9 0C 64-58. 0 C
QRHE 118MW 129MW 120MW
Dependent Variables
Tcool-Gas 303 0 C 307 0C 307 0 C
TFW-HRSG-in 297 0 C 301 0C 302 0 C
TComb-Gas-in 294 0 C 294 0 C 289 0 C
Flue gas flowrate in 1091kg/s 1,080kg/s 1076kg/s
HRSG
Condensed water 33.2kg/s 36.1kg/s 34.9kg/s
RHE
APHRSG 0.277bar 0.474bar 0.277bar
APpip, primary 0.059bar 0.070bar 0.044bar
APpzpe secondary 0.056bar 0.0690bar 0.038bar
mFCRec-Pri 278kg/s 259kg/s 258kg/s
mFG-Rec-Sec 705kg/s 690kg/s 687kg/s
TRHE-out 160 0 C 171 0 C 163 0 C
Table 3.6: Results for RHE flowsheet fuel flexibility. Area 'favoring' the design of
CoalB; AFWH1=6,728m 2 AFWH2=4,525m 2 . The changes in the values of rnFW Main
and PDeaerator are acceptably small, Section 3.4.2.

Variable RHE-B-1 RHE-A-2 RHE-A-3


Input Parameters
Fuel Flowrate 35.53kg/s 30kg/s 30kg/s
Slurry water flowrate 19.22kg/s 16.50kg/s 16.50kg/s
Atomizer Stream 2.96kg/s 2.50kg/s 2.50kg/s
flowrate
Air flowrate 332.5kg/s 311.2kg/s 311.2kg/s
Efficiency (LHV) 33.05% identical to col- 34.42%
umn 4
Independent and Key Dependent Variables
Pcomb,d 9.67bar 7.41bar 9.67bar
PComb, o 9.67bar 7.41bar
QComb 1MW 1MW
mFW Main 284 kg/s 289kg/s
PBLD1 74.3bar 74.3bar
rnBLD1 41.8kg/s 42.6kg/s
PBLD2 26.Obar 26.0bar
rnBLD2 19.5kg/s 20.9kg/s
PBLD3 11.5bar 11.5bar
rnBLD3 2.57kg/s 5.30kg/s
PDeaerator 11.3bar 10.6bar
QFWH1 89.9MW 91.7MW
QFWH2 51.0MW 55.0MW
TFG-HE-out
0
38-3 C 368 0C
QRHE 128MW 118MW
Dependent Variables
TCool-Gas 303 0 C 303 0 C
TFW-HRSG-in
0
297 C 297 0 C
TComb-Ga&-in 290 0C 288 0 C
Flue gas flowrate in 1072kg/s 1,091kg/s
HRSG
Condensed water 36.1kg/s 33.2kg/s
RHE
APHRSG 0.460bar 0.280bar
APpe primary 0.059bar 0.050bar
APpzpe secondary 0.056bar 0.038bar
mFG-rec-pri 258kg/s 276kg/s
mFG-re-sec 683kg/s 694kg/s
TRHFrout 170 0 C 160 0 C
Table 3.7: Results for DCSC flowsheet fuel flexibility. Area 'favoring' the design if
CoalA; AFWH1=6,158m 2 AFWH2=4,227m 2 . The changes in the values of rnFW,Main,
PDeaerator, and TnRW-Sepin are acceptably small, Section 3.4.2.

Variable DCSC-A-1 DCSC-B-2 DCSC-B-3


(no chiller (no chiller (chiller required)
required) required)
Input Parameters
Fuel Flowrate 30.00kg/s 35.53kg/s 35.53kg/s
Slurry water flowrate 16.50kg/s 19.22kg/s 19.22kg/s
Atomizer Stream 2.50kg/s 2.96kg/s 2.96kg/s
flowrate
Air flowrate 311.2kg/s 332.5kg/s 332.5kg/s
Objective Function 34.10% 32.48% 32.15%
(LHV)
Independent and Key Dependent Variables
PComb,d 12.8bar 16.9bar 12.8bar
PComb, o 12.8bar 16.9bar 12.8bar
OComb 1MW 1MW 1MW
mFW Main 286kg/s 274kg/s 279kg/s
PBLD1 65.7bar 57.9bar 65.7bar
rnBLD1 36.8kg/s 32.1kg/s 36.4kg/s
PBLD2 25.2bar 23.6bar 25.2bar
rnBLD2 17.0kg/s 13.1kg/s 25.2kg/s
PBLD3 14.7bar 15.3bar 14.7bar
rnBLD3 9.52kg/s 8.27kg/s Okg/s
PDeaerator 12.2bar 13.5bar 7.61bar
QFWH1 78.4MW 68.2MW 77.7MW
QFWH2 44.1MW 33.5MW 67.4MW
TFG-Sep-out 35.8 0 C 38.1 0 C 96.8 0 C
ODCSC-HX 118MW 127MW 116MW
mRW-Sep-in 189kg/s 186kg/s 218kg/s
Dependent Variables
TCool-Gas 292 0 C 283 0 C 292 0 C
TFW-HRSG-in
0
287 C 0
277 C 287 0 C
TComb-Gas-in 290 0C 287 0 C 298 0 C
Flue gas flowrate in 1,081kg/s 1,045kg/s 1,056kg/s
HRSG
Condensed Water 33.5kg/s 36.4kg/s 33.1kg/s
Sep..Col
APHRSG 0.865bar 1.53bar 0.86bar
APpipe primary 0.088bar 0.137bar 0.092bar
APpipe secondary 0.088bar 0.136bar 0.089bar
mFG-Rec-pri 276kg/s 255kg/s 255kg/s
mFG-Rec-sec 684kg/s 658kg/s 668kg/s
Rankine net power 396MW 420MW 403MW
TFW-Recov-out 158 0 C 165 0 C 1560C
argument applies when comparing Column5 of Table 3.9 to RHE-B-1.

Columns2&3 represent CoalA and CoalB respectively with the first set of area

specifications and prove the ideal flexibility when operating with RHE-AreaA. Columns

four and five prove ideal-flexibility when operating with RHE-AreaB. Since Columns2&4

show that the cycle is ideally flexible to the area changes while operating with CoalA,

and columns3&5 show the same while operating with CoalB, the process is ideally

flexible to FWHs' areas. More specifically, the area specifications do not affect the

choice of other design variables. This has positive implications to turbine manufactur-

ing, retrofitting, construction of plants with intent/uncertainty to upscaling etc. The

four columns which are introduced as belonging to two different hierarchical optimiza-

tion runs, one for each FWHs area, can be presented as a single larger hierarchical

optimization problem. The bigger problem has four objectives, which are the perfor-

mances of the different combinations of the two coals and the two area specifications,

instead of just the original two coals. The problem formulation is shown below where

a and b stand for the areas obtained from CoalA's initial design and CoalB's initial

design respectively.

Note that here too the only feasible solutions are ideally flexible designs within

the allowed tolerances ekL. Results in Table 3.9 show that 6 can be set essentially to

zero, i.e., the RHE process is ideally flexible to coal and FWHs areas variations. The

same behavior is encountered for the DCSC process but not shown here.

Behavior of Operation Variables

In this section, the optimal values of the operation variables are discussed for the

changes in the coal input. The values of the design variables essentially do not change

between different coal operations and thus are not discussed, yet they have a range

of near-optimal values. Unless otherwise specified, the discussion is relevant to the

118
ideally flexible design of Table 3.9.

Pressure, PComb In the RHE flowsheet, the optimum combustor operating pres-

sure for CoalB is higher than that of CoalA in order to allow for the additional thermal

recovery carried with the larger flue gas flowrates and larger amount of water vapor.

However, with larger operating pressure the pressure losses and the compensation

power requirements increase. Recall that the optimum pressure range for CoalA in

an RHE flowsheet is determined by two main factors, [24, 27]: (i) large amount of

water condensation and thermal recovery at the RHE, and (ii) flue gas compression

power requirements being close their minimum value. With a lower quality coal,

CoalB, the advantage of increasing the operating pressure for enhancing thermal re-

covery is diminished by the increase in the pressure losses and the compensation

power requirements; in scenarios where the coal quality is even lower, then the op-

timum pressure might be lower than that providing maximum thermal recovery in

order to avoid huge pressure losses.

When the operating pressure is limited by the design pressure as an upper bound,

a lower quality coal will suffer from inadequate recovery as seen in RHE-B-3 and

DCSC-B-3 of Tables 3.5&3.7. The lower recovery results in a lower temperature of

the feedwater entering the deaerator. Therefore, the deaerator pressure decreases. In

the above scenarios, the deaerator bleed is favored to be zero because its extraction

design pressure is now considerably larger than the deaerator pressure. The losses

also include the mixing of bleedsl&2 as they exit FWHs1&2 to a lower pressure in

the deaerator, and additional power requirements by the HP pump to elevate the

feedwater pressure over a larger range. Therefore, the design of the process should

take into consideration the different coals involved and simultaneous optimization

should be performed to obtain the ideally flexible designs like in Tables 3.6& 3.8 and

particularly Table 3.9.

119
Combustor Duty, QComb For both processes and both coals, the combustor duty

again takes on the minimum allowed value, 1MW, as expected. It is preferable

to transfer thermal energy to the high-temperature section rather than to the low-

temperature section of the cycle.

Thermal Recovery, QRHE/QDCSC-HX Utilizing CoalB, the flue gas contains a

larger amount of water vapor due to the increased flowrate of slurry water, atomizer

stream, and additional water content in the higher coal flowrate. The higher water

content results in a larger heat recovery at the RHE/DCSC. However, the additional

recovery at the low temperature section is at the expense of the heat transfer at the

high temperature section; transferring more thermal energy through the recovery unit

signifies transferring less thermal energy at the HRSG.

For the RHE flowsheet, the recovery section independent variable is chosen to be

the temperature of the flue gas exiting the RHE, TFG-RHE-Out. During the different

operations, the value of TFG-RHEout does not change because by increasing the oper-

ation pressure the pinch point of the surface heat exchanger is reached at the outlet

of the flue gas, yet the recovered duty changes due to the different constituents of the

flue gas. It is worth mentioning that the RHE experiences a double pinch, i.e., the

MITA is encountered at two locations, at optimum conditions. As operating pres-

sure increases above atmospheric, recovery rapidly increases predominantly due to

the decreasing temperature of the flue gas leaving the RHE till the pinch is reached

at that outlet as opposed to just occurring at the dew point, [24, 27]. Any further

increase in flue gas pressure results in no change in the outlet temperature and in

very small increase in recovery and water condensation due to the increase in the

partial pressure of the water vapor. The pressure losses increase accompanied with

the operating pressure increase dominates over the insignificant additional recovery.

Therefore, optimization results in a double pinch where a higher pressure operation

120
suffers from the dominating increase in pressure losses and a lower pressure operation

suffers from significantly lower recovery.

For the DCSC process, the two liquid streams of the DCSC-HX are enforced to be

balanced, as described in Section 3.4.4. The hot end is governed by the temperature of

the recirculating water entering the DCSC-HX/leaving the separation column which

is identical to the temperature of the bottom stage of the column. The separation

column's bottom stage is the first site of condensation in the DCSC recovery unit and

therefore analogous to the point of water condensation of the flue gas in the RHE.

Main Feedwater Flowrate, ThFW, Main The amount of thermal energy absorbed

by the increased flow of water entering the combustor (slurry water, moisture in coal,

atomizer stream) increases. Most of this thermal energy is transferred to the Rankine

cycle at the RHE/DCSC during condensation instead of being transferred at the

HRSG. rhFW, Main decreases due to the decrease in the thermal energy available at

the HRSG, reducing the flow through the turbines and decreasing the power output.

The decrease in mFW, Main is analogous to results obtained in [24, 27] where water

addition into the cycle is preferred to be smaller. The oxidizer's increased flowrate also

results in a similar behavior but to a smaller extent than the increased water flowrate.

The main stream flowrate is not significantly affected because of the compression

enthalpy rise (CER) of the flue gas. The CER is the enthalpy added to the flue gas

by the compensation power requirement (CPR) of the recycling fans, [24, 27]. This

increase in enthalpy is eventually transferred to the Rankine cycle which contributes

in obtaining a working fluid flowrate close to nominal, and results in a comparable

gross power output.

Deaerator and Deaerator Bleed, PDeaerator and PBLD3 and mBLD3 When the

design pressure is lower than the optimal pressure required for the process, recovery

suffers and the deaerator pressure decreases. However, with lower-quality coals, when

121
the design pressure allows reaching the optimal operation pressure, the increase in

thermal energy transfer at the RHE/DCSC and the lower feedwater flowrate results

in a higher temperature of the feedwater entering the deaerator. Therefore, a smaller

bleed flowrate is required to cause saturation at the deaerator tank. The lower bleed

flowrate results in a smaller pressure drop through the connection pipes, allowing for

a larger deaerator pressure.

FWH Bleeds Flowrate and FWHs Duty Transfer, rnBLD(1&2) and QFWH(1&2)

Lower feedwater flowrate requires smaller amounts of heating from the FWHs, con-

trolled by the bleeds' flowrates and FWHs' duty transfer. Therefore, the optimum

operation of CoalB requires lower bleedl&2 flowrates and smaller QFWLH(1&2)-

DCSC recirculating water flowrate, rhRW-Sep-in In the DCSC, changing from

CoalA to CoalB reduces the main feedwater flowrate trough the Rankine cycle, rnMain,

as explained above, which signifies a decrease in the working fluid flowrate through the

recovery unit. According to the optimality of the active constraint, balanced DCSC-

HX, the flowrate of the recirculating stream in the DCSC-HX/exiting the separation

column decreases. A smaller amount of recirculating water exiting the column is

required and a larger amount of condensed water from the flue gas within the column

means that the amount of the recirculating water entering the separation column,

rnRW-Sep-in, decreases.

Discussion for Efficiency and Rankine cycle Power While the heat rate

burned/consumed is constant, the gross and net power outputs are not. Using a

lower heating value coal results in a decrease in the exergy of the heat input because

more thermal energy is transferred at the low-temperature section of the cycle rather

than the high-temperature section due to the larger flows, particularly water flow,

into the combustion process, as detailed in [24, 271. The larger water flow here is due

122
to the coal's larger moisture content and flow, larger water slurry flow, and larger

atomizer stream flow. In addition, the atomizer stream is withdrawn from the steam

expansion line contributing to further decrease the gross power output. Coupled with

larger power requirements for the larger amount of air flow input and flue gas to be

purified, the overall efficiency decreases with lower quality coals despite the presence

of thermal recovery. However, the decrease in the gross power is not large. The reason

is that the compression requirements cause the flue gas to acquire larger amounts of

enthalpy which are eventually transferred to the Rankine cycle to produce additional

gross work; but as detailed in [24, 27], the increase in the compression power require-

ments are larger than the extra work they contribute in producing which results in

an overall decrease in efficiency.

3.5.3 Flexibility to Input Flows and Parameters, (Air Flow,

Slurry water Flow, atomizer Stream Flow, and Oxi-

dizer Stream Oxygen Purity)

In the above study, input parameters like the air flowrate, slurry water flowrate,

and atomizer stream flowrate are proportional to the coal flowrate as determined

by technological constraints. It is plausible that some of these specifications are

revisited after the plant is built, in particular for technologies under development,

such as pressurized OCC. As seen above, increasing the coal flowrate when changing

from CoalA to CoalB increases each of those parameters' flowrates and results in a

total decrease of exergy of the combustion gas. Notice that the increase in flow of

each one of those parameters contributes in decreasing the exergy in different amounts

but in the identical manner: more thermal energy transferred at the low temperature

section at the expense of the thermal energy transferred at the high temperature

section. Now since the processes are able to handle the summed large decrease in

123
exergy, it is possible that they are also able to handle a decrease in exergy due to an

independent increase of one or more of these flowrate. The same flexibility applies

regarding the decrease in flowrate of those parameters due to decreasing the coal

flowrate when changing from CoalB to CoalA. Therefore, the processes are expected

to be ideally flexible to changes in the input parameters of air flowrate, slurry water

flowrate, and atomizer stream flowrate. For brevity, the results of this flexibility

assessment are not shown here. Of course changes in flowrates must fall within the

acceptable operation ranges of the combustor, herein modeled approximately; also it

is necessary that all the utilized coal is oxidized.

The flexibility of the processes to the air flowrate has yet another positive implica-

tion regarding the ASU's oxygen purity. It is optimal for the process to operate with

the smallest possible oxidizer and air flowrates, given that all the fuel is oxidized, be-

cause it requires the least compression requirements throughout the process; therefore,

a change in purity might require a change in the input air flow. The oxygen purity

plays an important role in determining the ASU's and CSU's power requirements,

which are to a large extent independent from the energy conversion process and the

power delivered by the power cycle, but the process is considered ideally flexible to

the oxygen purity as shown next. Consider, for example, a given flowrate and purity

of the oxidizer stream that is capable of oxidizing a fixed amount of fuel. Chang-

ing the purity of the oxidizer stream while keeping the same oxidizer flowrate and

pressure, results in the same flowrate of the flue gas but a different oxygen content;

argon and nitrogen content may change too but their concentrations are insignificant

to begin with. The effects of the change in the oxygen content in the flue gas is

insignificant; first, the range of variation of the oxygen purity in the oxidizer stream

is relatively small, within 85 to 98 vol%, [46], and even smaller after mixing with the

flowrate of the primary recycling stream, which is significantly larger than that of

the oxygen stream; second, CO 2 and then H 2 0 are the predominant species of the

124
flue gas and the recycling streams, and these species are responsible for dictating the

streams' properties particularly the thermal capacity. Even if the thermal capacity

of the flue gas is slightly affected, the effect on the power cycle is again insignificant

because much larger changes in flue gas properties and recycling flowrates, seen in the

coal flexibility, are easily accommodated. The acid condensation temperature may

change, with concerns towards an increase in acid dew point and condensation prior

to recovery; however, the optimum temperature of flue gas heading to the recovery

unit is already significantly higher than the acid dew point, hence, acid condensation

is not an issue. Also, the water condensation in the recovery unit is not affected

because the H 2 0 concentration is not significantly changed. As a result, given com-

plete oxidation of the coal, the processes are not affected by a change in the oxidizer

stream purity for the same flowrate. Meanwhile, from the air flow flexibility above,

the processes are ideally flexible to changes in the oxidizer flowrate while maintaining

the same purity. As a conclusion, the processes are flexible to a general change in the

purity of the oxidizer stream, even when accompanied by a change in its flowrate, as

long as the fuel is completely oxidized.

In other words, the processes are flexible to the air flowrate whether the change

in the flowrate is due to a change in the coal flowrate, the preferred ratios of oxygen-

to-coal, or due to a change in the oxygen purity of the ASU (or all three). A total

oxidation of the fuel is assumed in the above discussion which is trivially the most

efficient way to operate. It is true that the efficiency will change with different values

of oxygen purity, but these changes can be accommodated by the flexible design during

operation to achieve and maintain the maximum possible performance for any value

of oxygen purity. Here again any change should be within the acceptable operation

region of the combustor which is not modeled here in component level details.

125
3.5.4 Flexibility to Ambient Conditions

In the processes considered the condenser rejects thermal energy to cooling water from

a river or reservoir. Pressure and humidity changes of the ambient have no effect on

the condenser's operation; the condenser's operating pressure is defined by the type

of the working fluid and the condenser's temperature. Typically, the condenser's

pressure is substantially lower than the atmospheric pressure. Also, ambient pressure

and humidity do not affect the deaerator.

However, the ambient temperature, more specifically the cooling water tempera-

ture, creates uncertainties to the processes because it affects the condenser's operating

temperature. As a result the condenser's operating pressure and the temperature of

the feedwater entering the recovery unit changes, which therefore changes the recov-

ered thermal energy, the deaerator operation, and so on, as well as the temperature of

the flue gas entering the CSU and the CSU power requirements. To assess the flexibil-

ity towards the temperature of the cooling water, a similar methodology is followed as

the one presented above for the coal. Note that the temperature of the atmospheric

air slightly affects the ASU process, however, these effects do not propagate to the

power cycle since the oxygen pure stream is delivered by an intercooled compressor

with intercooling temperature of 60C, which is well above the ambient temperatures.

The power of the ASU can be slightly altered with ambient temperature variations,

but this is not represented here due to the limited modeling of the ASU. The initial

condenser temperature and pressure are 29.73'C and 0.042bar. A temperature rise

of 10'C, which is representative of seasonal and location variations, is evaluated re-

sulting in a condenser pressure of 0.073bar. For brevity, only the RHE process with

CoalA and AreasA is presented. Hierarchical optimization for changes in ambient

temperature is performed with the initial guess being that of the ideally flexible de-

sign of the above hierarchical optimization represented in Table 3.9. Optimization

results in no change in the values of the design variables even with large tolerances,

126
E, which means that the optimal design for the high ambient temperature is identical

to that of the low ambient temperature; the ideal flexible design for the changes in

coal, input streams flows and parameters, and FWHs Areas is simultaneously ideally

flexible to ambient conditions. Other optimization runs with different initial guesses

were performed to make sure that the claimed ideally flexible design is not a local

optimal for the new high ambient temperature condition. Results of the temperature

flexibility are shown in Table 3.10.

Response to Variations in Ambient Conditions

It is obvious that the efficiency decreases with higher ambient temperature due to

the higher condenser pressure. Moreover, with larger ambient temperatures, the

minimum value of the temperature of the flue gas exiting the recovery unit is larger

due to the higher temperature of the feedwater entering the RHE; this results in lower

recovery and larger CSU power requirements. The optimum design of the process

with a high ambient temperature is identical to that of the low ambient temperature

proving the ideal flexibility to ambient conditions, but optimization of the operation

variables is required to ensure the maximum performance with the changes in ambient

conditions.

The operating pressure increases in order to allow for a larger thermal energy

transfer at the RHE by avoiding the limitations of the pinch at the onset of flue

gas condensation. With the initial operation pressure, the higher temperature of

the feedwater entering the RHE results in having the pinch at the onset of water

condensation and not at the cold end of the RHE (Hot-stream-out/Cold-stream-in),

and limits the amount of condensed water and the amount of recovered thermal

energy. Increasing the operating pressure allows the pinch to occur at both the point

of flue gas water condensation and at the flue gas outlet allowing for lower flue gas

exit temperature and more water condensation. Upon optimizing the process with

127
the high ambient temperature condition, the amount of water condensed reaches

33.Okg/s, slightly less than the maximum amount possible with the lower ambient

temperature conditions.

Notice that with larger operating pressures, the pressure losses increase. Further

increase in the ambient temperature may favor increasing the pressure to values that

do not acquire maximum recovery in order to save on pressure losses and compensation

power requirements.

It is worth mentioning that nine out of the 10 operation variables for the RHE

cycle, and 10 out of the 11 for the DCSC cycle are governed by active constraints.

The combustor's operating pressure is the remaining operation variable that is not

governed by an optimality criterion. Attempting to optimize for all the operation

variables instead of following the proven criteria, [24, 27, 38, 28, 29] and Chapters 1&2,

is likely to result in suboptimal solutions which do not satisfy ideal flexibility.

3.6 Conclusion

The operation of a powerplant is subject to several uncertainties in the inputs or

parameters or surrounding conditions. A realistic operation of a coal power plant

involves the utilization of different types of coals depending on several unpredictable

factors like market, environment, supply, demand, etc. Slurry water, atomizer stream,

and air flow requirements among other parameters are also subject to change. More-

over, changes in ambient conditions can significantly alter the power generation pro-

cess if not properly accounted for. For profitable power generation the plant should

accommodate the different possible uncontrollable conditions with high production

efficiency.

Some process variables are design variables that cannot change during operation,

while operation variables can. Herein, using relatively detailed models, optimization

128
of two OCC processes is performed while facing uncertainty in the input parameters,
in particular coal types and specifications and ambient temperature. Uncertainty in

input streams flowrates and in oxygen purity of the oxidizer stream are also discussed.

Results show that the studied concept of pressurized OCC is ideally flexible. More

specifically, changes in inputs can be accommodated without any compromise in the

processes' performance compared to a process specifically designed for the new values

of the inputs. The study concludes that stochastic optimization is not needed for

designing the flexible powerplant. In essence, the uncertainties in input conditions

and parameters need not to be quantified but merely the range of input conditions

needs to be taken into account during design. The ideal flexibility of the processes is

discovered with the help of the optimum criteria of operation presented in [28, 29, 381

and Chapters 1&2. These criteria allow finding the ideal flexible design, during design,

and provide straight forward guidelines, during operation.

It is particularly interesting and convenient that changes in the input specifica-

tions, which require changes to the operation variables, can be accommodated with

minimal efforts. In essence, after finding the ideally flexible design, a change in im-

posed conditions, even as large as using a significantly different coal, FWHs areas, and

ambient temperatures, does not require any additional modeling and optimization ef-

forts to maintain optimum operation, but rather only a change in variables to satisfy

the relevant criteria. The same applies to other operating parameters like slurry water

flowrate, air flowrate, atomizer stream flowrate, and oxidizer stream oxygen purity.

Results also show insensitivity regarding the design variables near their optimum.

In particular, the pressurized OCC process is found to be ideally flexible regarding

the FWHs area specifications, which are also a part of the plant's design. The insen-

sitivity towards design variables is initially observed in [28, 29] and Chapter 1. These

results strengthen the possibilities and interest in retrofitting existing powerplants.

The fixed designs of an operating powerplant, in particular fixed heat exchanger areas

129
and fixed turbine expansion line, are likely to achieve high performance as long as they

are not too far away from optimum. Also their performance will not suffer with the

unpredictable changes in the input specifications and parameters of operation. Addi-

tional positive implications are seen regarding turbine manufacturers when designing

the expansion line and extraction locations which are shown to be less contingent on

the other power plant specifications like FWHs areas.

When changing the coal, the fuel heat rate input is held fixed instead of the

net output power for three reasons. The main motivation is that the processes are

limited by the combustor's and HRSG's sizes and thermal loads. An additional

motivation is that this allows to keep the same basecase reference conditions of the

cycle (HRSG and recycling pipes' reference/basecase pressure losses, turbine and heat

exchanger's sizing etc.). Third, this allows to maintain the working fluid flowrate close

to nominal; this achieves high performance for the turbines and also eliminates the

need for performance curves. If in contrast, the cycle operates under fixed net work

output, it has to compensate for the additional compression requirements caused

by using a lower quality coal, CoalB; i.e., the Rankine cycle is required to produce

more gross power output. Larger gross power output requires even more fuel and

inlet stream flows and result in even more compression requirements that need to

be compensated for. As a result, the feedwater flowrate through the HRSG and

turbine expansion line increases, and will result in significant deviation from design

conditions. Additionally, keeping the fuel flowrate constant when using a lower quality

coal, which forms a flue gas with lower exergy, results in lower thermal energy transfer

at the HRSG; as a result, the feedwater flowrate capable of flowing through the HRSG,

and eventually through the expansion line, will decrease, causing deviation from the

expansion line design conditions.

It is noteworthy to discuss the flexibility presented herein for the pressurized

OCC in regards to regular and pressurized coal processes. In standard, non OCC,

130
coal power plants, thermal energy is transferred from the flue gas to the working fluid

only at the boiler due to the absence of the low temperature heat transfer section, a

RHE or a DCSC. Even if a recovery section is incorporated, if the flue gas pressure

is atmospheric then an insignificant amount of thermal recovery is obtained; for a

pressurized coal process without carbon capture the compression enthalpy rise of the

high rates of recycling in the pressurized OCC process is absent. Therefore, operating

at the same boiler load, the amount of thermal energy capable of being transferred to

the working fluid decreases. As a result, the working fluid flowrates are likely to be

significantly affected, altering the turbine expansion line characteristics; the response

of the expansion line might not be suitable with the cycle's original design, extraction

positions and pressures, FWHs sizes, reheat stream pressure and temperature etc.

Therefore, non-OCC coal power generation is expected to be less flexible than the

processes considered herein; however, a detailed investigation is warranted to check

this claim.

3.7 Future Work

It is highly possible that the processes are flexible to a varying load as motivated next.

The processes are flexible to flow changes for the same heat rate input, and capable of

attaining the different maximum efficiencies for the different parameters imposed; the

performance change due to the change in compression requirements and to the change

in the exergy of the flue gas; the two factors are to a large extent independent. Now

assume a CoalC with specifications in between those of CoalA and CoalB. Operating

at the same load and input specifications ratios, the input streams of CoalC and flue

gas flowrates are larger than those of CoalA but lower than those of CoalB. The exergy

of the thermal energy of the flue gas resulting from CoalC operation is lower than

that of CoalA but higher than that of CoalB. Now consider a decrease in the flowrate

131
of the CoalC in order to operate at a lower load. The streams flowrate will get closer

to the base load of CoalA, but the heatrate decreases and thus the availability of the

thermal energy will get closer to that of the base load of CoalB; such an operation

is analogous to having CoalB with base load operation but with lower input streams

flowrates, which can be accommodated by the flexible design, Section 3.5.3. The same

argument can be applied for an increase in flowrate of CoalC, considering that the

thermal load of the combustor and HRSG are not limiting. Therefore, the processes

might very well be flexible to varying load. However, load changes might be larger

than the range allocated by the two coals and that is why a detailed study is intended

for future work.

Part-load, presented in Chapter 4, is another important topic but requires an

expansion line model accounting for off-design operations. The task is to determine

how to operate optimally for part-load and more generally how to operate optimally

facing the different and uncertain load requirements. It is intriguing to consider if

the process is also ideally flexible to load changes. If not, then does the design of a

process require knowledge of the distribution of the expected load operations or just

the range of the part-load? The next chapter handles a detailed study of the process

facing variations and uncertainty in the thermal load. Using optimization the process

is designed to be flexible for the load.

132
Table 3.8: Results for DCSC flowsheet fuel flexibility. Area 'favoring' the design of
CoalB; AwrH=5,552m 2 AFWH2=3,561m 2 . The changes in the values of rnFW,Main,
PDeaerator, and rnRW-Sep-in are acceptably small, Section 3.4.2.

Variable DCSC-B-1 DCSC-A-2 DCSC-A-3


(chiller required) (no chiller (no chiller
required) required)
Input Parameters
Fuel Flowrate 35.53kg/s 30.00kg/s 30.00kg/s
Slurry water flowrate 19.22kg/s 16.50kg/s 16.50kg/s
Atomizer Stream 2.96kg/s 2.50kg/s 2.50kg/s
flowrate
Air flowrate 332.5kg/s 311.2kg/s 311.2kg/s
Objective Function 32.43% 34.07% 34.05%
(LHV)
Independent and Key Dependent Variables
PComb,d 15.5bar 12.8bar 15.5bar
PComb, o 15.5bar 12.8bar 12.5bar
QComb 1MW 1MW 1MW
mFW Main 274kg/s 285kg/s 280kg/s
PBLD1 56.0bar 65.Obar 56.Obar
rnBLD1 30.9kg/s 36.4kg/s 31.5kg/s
PBLD2 23.3bar 25.2bar 23.3bar
rnBLD2 12.7kg/s 16.9kg/s 15.4kg/s
PBLD3 15.4bar 16.7bar 15.4bar
rnBLD3 9.07kg/s 9.12kg/s 9.56kg/s
PDeaerator 13.3bar 12.4kg/s 14.1bar
QFWH1 65.5MW 77.7MW 66.8MW
QFWH2 32.4MW 43.9MW 39.8MW
TFG-Sep-out 61.1 0 C 35.0 0 C 35.0 0 C
QDCSC-HX 124MW 118MW 115MW
mRW-Sep-in 186kg/s 189kg/s 189kg/s
Dependent Variables
Tcool-Gas 280 0C 291 0C 280 0 C
TFW-HRSG-in 0
274 C 285 0 C 274 0 C
TComb-Gas-in 283 0C 288 0C 278 0 C
Flue gas flowrate in 1,039kg/s 1,078kg/s 1,059kg/s
HRSG
Condensed Water 36.0kg/s 33.4kg/s 33.3kg/s
SepCol
APHRSG 1.30bar 0.86bar 0.84bar
APpipe primary 0.124bar 0.088bar 0.087bar
APpipe secondary 0. 123bar 0.088bar 0.089bar
mFG-Rec-pri 254kg/s 276kg/s 274kg/s
mFG-Rec-sec 653kg/s 682kg/s 665kg/s
Rankine cycle power 417MW 423MW 422MW
TFW-Recov-out 162 0C 155 0C 154 0 C
SteplA StepiB Step2
Objective maxX,YA ?7A(X, YA) maxX,YBq B(x, YB) maxX,YA,YB (x, YA, YB) i =
A or B
Constraints gA(x,yA) < gB(x,yB) < 0 gA(x,yA) 0 & gB(X,yB)
0 & 7A(X,YA) AA*-A &
rB (X,YB) 7B* EB
Desired x4,y~,?A(X*,Y*)=A* x* YB
*7B(XIBYB 77B xIYIYB,I7A (X)I = A
Results ?7^, and 7B* B* I=
(assuming B
feasibility)

SteplAa StepiBa SteplAb SteplBb


Objective maxX,YA,. 7A,a(X, YA,a) maxx,YB,. ?7B,a(X, YB,a) maxx,YAb rJA,b(X, YA,b) maxX,YBb ?7B,b(X, YB,b)
Constraints gA,a(x, YA,a) < 0 gB,a(X, YB,a) 0 9A,b(x, YA,b) 0 9B,b(X, YB,b) 0
Desired A,,y*,a BaYB,a XA,b,YA,b *B
,B,b
Results
(assuming
feasibility)
(A B~a) X _ Ba) 7 A,b)*7A 'q(*b X b) (Bb)
77
?7A,a(XAa, YA,a) = ra r7B,aBB,a, yB,a) = Ba Ab( Ab, YA,b)
= bbBb ?B,b( B,b, XB,b)
B bb
Step2
Objective maxXYAaYBabYBb rlij(X, YA,a, YB,a, YA,b, YB,b) i = A or B and j = a or b
Constraints gk,L(x,yk,I) 0
77k,1(X, Yk,) k,) -- Ek,l
V k E{AB} and 1 E{ab}
Desired x Y,a, YB**a, Y YB*,bbA*b
Results
(assuming
feasibility)
?7A,a(X, YAa) =IAa ** (Aab)*
1
TB,a ( X, yBa) = BPa )a
7
7A,b , y~b T|b b)*
and 77B,b(x 7
, Bb 7Bb (Bb)
Table 3.9: Fuel and Area flexibility for the RHE process after hierarchic optimization. The
process is designed to be ideally flexible under fuel uncertainty while operating with the
first pair of area specifications (AFWH1 = 7, 069m 2 and AFWH2 = 5, 254m 2 ), columns2&3.
It is also ideally flexible to the fuel uncertainty while operating under the second pair of
area specifications (AFWH1 = 6, 728m 2 and AFWH2 = 4, 525m 2 ), columns4&5. The changes
in the values of TiFW,Main and PDeaerator are acceptably small, Section 3.4.2.

Area specifications Areas favor CoalA design Areas favor CoalB design
Variable RHE-A new RHE-B new RHE-A new RHE-B new
variables variables variables variables
Input Parameters Input Parameters
Fuel Flowrate 30kg/s 35.53kg/s 30kg/s 35.53kg/s
Slurry water flowrate 16.50kg/s 19.22kg/s 16.50kg/s 19.22kg/s
Atomizer Stream 2.50kg/s 2.96kg/s 2.50kg/s 2.96kg/s
flowrate
Air flowrate 311.2kg/s 332.5kg/s 311.2kg/s 332.5kg/s
Efficiency (based on 34.42% 33.07% 34.41% 33.08%
LHV)
Independent and Key and Varaibles
PComb, d 9.67bar 9.67bar 9.67bar 9.67bar
PComb, o 7.41bar 9.67bar 7.41bar 9.67bar
QComb 1MW 1MW 1MW 1MW
mFW Main 306kg/s 301kg/s 306kg/s 303kg/s
PBLD1 99.Obar 99.Obar 99.Obar 99.Obar
rnBLD1 62.2kg/s 61.4kg/s 62.0kg/s 61.5kg/s
PBLD2 26.0bar 26.Obar 26.Obar 26.Obar
rnBLD2 14.7kg/s 12.8kg/s 16.7kg/s 15.2kg/s
PBLD3 16.8bar 16.8bar 16.8bar 16.8bar
rnBLD3 9.27kg/s 6.77kg/s 7.11kg/s 4.57kg/s
PDeaerator 14.7bar 15.7bar 13.4bar 14.2bar
QFWH1 138MW 136MW 137MW 136MW
QFWH2 377MW 32.7MW 43.2MW 39.2MW
TFG-RHE-out 36.9 0 C 38.1 0 C 36.9 0 C 36-9 0 C
QRHE 122MW 133MW 122MW 133MW
Dependent Variables
Tcool-Gas 327 0 C 328 0 C 327 0 C 327 0 C
TFW-HRSG-in 322 0 C 322 0 C 321 0 C 321 0 C
TComb-Gas-in 309 0 C 311 0 C 309 0 C 311 0 C
Flue gas fkiwrate in 1,138kg/s 1,118kg/s 1,137kg/s 1,125kg/s
HRSG
Condensed water 33.2kg/s 36.1kg/s 33.2kg/s 36.1kg/s
RHE
AXPHRSG 0.265bar 0.437bar 0.266bar 0.448bar
APpi, Primary 0.054bar 0.055bar 0.054bar 0.055bar
APpi, secondary 0.035bar 0.051bar 0.035bar 0.051bar
mFG-Rec-pri 281kg/s 262kg/s 281kg/s 265kg/s
FG-Rec-sec 737kg/s 724kg/s 736kg/s 729kg/s
Rankine cycle power 409MW 405MW 409MW 408MW
TFW-Recov-out 164 0 C 175 0 C 163 0 C 174 0 C
Table 3.10: Hierarchical optimization results for the RHE process with tem-
2
perature uncertainty. FWHs' areas are RHE-AreaA: AFWHi = 7,069m and
AFWH2 = 2, 524m 2 . The designed for the ideally flexible process facing coal,
areas, and flows uncertainty is also ideally flexible to the temperature uncer-
tainty. The changes in the values of TmFW,Main and PDeaerator are acceptably
small, Section 3.4.2.

Variable Low Tab Design High Tamb Design = op-


(RHE-A Column2 of timization of operation
Table 3.9) variables of Low Tamb
Design
Input Parameters
Fuel Flowrate 30kg/s 30kg/s
Slurry water flowrate 16.50kg/s 16.50kg/s
Atomizer Stream 2.50kg/s 2.50kg/s
flowrate
Air flowrate 311.2kg/s 311.2kg/s
Tcondenser
0
29.73 C 39.73 0 C
Pcondenser 0.042bar 0.073bar
Independent and Key Dependent Variables
Efficiency (LHV) 34.34% 33.23%
PComb, d max {PComb, oLow Tamb, PComb, 0 lligh Tamb} 8.92bar
PComb, o 7.41bar 8.92bar
QComb 1MW 1MW
rhFW,Main 306.0kg/s 307.0kg/s
PBLD1 99.Obar 99.Obar
ThBLD1 62.2kg/s 62.5kg/s
PBLD2 26.0Obar 26.Obar
rhBLD2 14.7kg/s 13.9kg/s
PBLD3 16.8bar 16.8bar
rhBLD3 9.27kg/s 8.15kg/s
PDeaerator 14.7bar 15.2bar
QFWH1 138MW 139MW
QFWH2 37.7MW 35.6MW
TFG-RHFrout 36.9 0 C 46.9 0 C
QRHE 122MW 120MW
Dependent Variables
TCJool-Gas 327 0 C 327 0C
TFW-HRSG-in 322 0C 322 0C
TComb-Gas-in 309 0C 312 0 C
Flue gas flowrate in 1,138kg/s 1,141kg/s
HRSG
Condensed Water RHE 33.2kg/s 33.0kg/s
APHRSG 0.265bar 0.385bar
APpipe primary 0.054bar 0.048bar
APpipe secondary 0.035bar 0.046bar
mFG-Rec-pri 281kg/s 282kg/s
mFG-Rec-sec 737kg/s 739kg/s
Rankine cycle power 409MW 401MW
0 1690C
TFW-Recov-out 164 C
Chapter 4

Pressurized OCC Process Ideally

Flexible to the Thermal Load

4.1 Summary

Pressurized oxycoal combustion process is optimized for variable thermal loading

(100% to 30%). The steam expansion line behavior is accurately represented based

on manufacturer data. Simulations with the nominal design and nominal operation

are then performed with variable loads to determine the level of performance de-

crease if no optimization is performed. Finally, optimization of operation for a fixed

design, and simultaneous optimization of design and operation are performed. The

design optimization for a specific load does not include the redesign of the turbines

to a process specifically designed for this load. However, the design variables of the

turbine expansion line, namely the extraction bleeds, are considered. At each load,

the performance of the process designed for nominal load is compared to the maxi-

mum possible performance obtained when designing the process for that specific load.

Thanks to the thermal recovery section, the process exhibits ideal flexibility to load

variations (not accounting for efficiency variations in the air separation unit, which

137
is accurate if oxygen storage is assumed), unlike Rankine cycles without pressurized

recovery. Consequently, there is no need to optimize for an expected distribution of

load operations. Finally, the process maintains supercritical operating conditions (no

phase change of FW in the HRSG) over larger ranges of thermal loads.

4.2 Motivation

Another very important disturbance to the process is variable loading, and is consid-

ered herein. At the time of plant design, the loading in uncertain. Load variations as

well as uncertainty in load is increasing in importance, [47, 481, especially with the rise

of the renewable but intermittent electric energy production, particularly wind and

solar. Therefore, because the process in not always operating at nominal conditions,

it is crucial to design the process for an overall maximum performance, rather than

the maximum performance at nominal conditions; in general, this is very challenging

and exhibits strong tradeoffs.

Studies regarding the performance of power generation processes at partload are

available in literature particularly for gas cycles and combined cycles, where significant

achievements in the design of the gas turbines allow them to operate flexibly to load

variations, [49, 50]. Fewer studies of partload are available for Rankine cycles or

the bottoming cycle of a combined cycle. In [47] a combined cycle is addressed to

obtain high performance at partload and meet the emissions criteria because of the

efficiency decrease with the decrease in operation load; the addressed parameters for

optimization are those related to the gas process where the gas turbines are fitted

with guiding vanes and preheating of the gas cycle air improve the overall efficiency

at partload. However, the parameters of the Rankine cycle itself are not addressed.

Similarly pertaining to the gas cycle section of the combined cycle, [511 addresses

the importance of guiding vanes and [52] studies different types of commercial gas

138
turbines, while maintaining the conditions for the bottoming Rankine cycle at nominal

in order to prevent alterations in its behavior. Solar thermal power generation and

organic Rankine cycles are also studied under the influence of partload, [531, therein

however, the Rankine cycle design and operation are not considered but rather the

size of the solar field for the minimum levelized cost of electricity production for

a given load schedule. Judes and Tsatsaronis [54] presents a comprehensive study

of a simplified single pressure combined cycle power plant under variable loading;

there too the main variables are the choice of the gas turbines and the temperature

approach of the heat recovery steam generators. The Rankine cycle is a simple two-

stage turbine with no reheat or regeneration. As will be demonstrated in this study,

regeneration and thermal recovery of a complex Rankine cycle have an intertwining

and complex behavior and requires a dedicated evaluation. Other studies involving

the partload flexibility of the Rankine cycle deal with comparing the valve throttling

method to sliding pressure boilers method for partload operations, or even hybrid

methods combining both, [55, 56, 57]. However, the optimal design and operation of

the process are not addressed as herein.

Herein, the pressurized oxycoal combustion process presented in Chapters 2&3,

seen in Figure 4-1, is optimized for variable thermal loading. The first challenge in the

analysis of load variations is determining the steam turbine expansion line behavior

as the operation deviates from the nominal conditions. Experimental isentropic effi-

ciency data are first analyzed in Section 4.3 to determine the expansion line behavior

and performance curves. Then, Section 4.4 presents the modeling approach. The

change of load also requires changes in several parameters to satisfy the operating

requirements of the units. For instance, the requirement on excess oxygen in combus-

tion implies that the oxygen stream is proportional to the coal stream. Moreover, for

realistic representation of the process, the pressure losses in the heat recovery steam

generator (HRSG) and in the recycling pipes are evaluated and implemented. Then,

139
the optimization formulation is presented in Section 4.5, i.e., optimizing the process at

different loads (100% to 30% with 10% increments) taking care in modeling the design

variables particularly the bleeds' extraction pressures. Following proven thermody-

namic criteria of optimum operation, [24, 27, 28, 29, 38] and Chapters 1&2, active

constraint optimization is utilized to simplify the optimization problem and avoid

convergence to suboptimal local optima. Finally, results are presented in Section 4.6

where the importance of optimization is reflected by the significant performance im-

provement of the variable load operations. Moreover, the flexibility of the process to

partload is evaluated by comparing the performance of a fixed design over the range

of loads to the maximum possible performance; the latter is the performance of the

processes designed for each of the specific loads. In designing for a specific load, the

turbines are considered to allow for the whole range of load operations, thus having

the nominal design, as justified in Section 4.5. Results show that due to the recovery

section, the process is ideally flexible to variable load, i.e., a fixed design operating at

variable load matches the maximum possible performance for each load. In contrast,

Rankine cycles without pressurized recovery do not share this favorable property.

Other important benefits of the recovery section are explained.

4.3 Turbine Performance Curves

The thermal load (TLoad) is the ratio of the coal flowrate into the combustor to

that of the full load nominal operation (TLoad= rcoal, actual


m'koal, nominal
). As the thermal load

changes, the amount of thermal energy transported from the flue gas to the working

fluid of the Rankine cycle changes, thus changing the behavior of the working fluid.

The deviation of the turbine operation from the nominal conditions changes their

efficiency and power output. Therefore, an accurate representation of the expansion

line versus the relevant variations in the process is required. Stodala, [58], is credited

140
for one of the earliest attempts to assess the flow variations of a multistage turbine

by developing the ellipse law using experimental data. The rule is an experimentally

derived equation that relates the steam mass flowrate, inlet and outlet pressures, and

inlet temperature at off-design modes. Here however, the efficiency and power output

of the turbine are also required. Therefore, the off-design operation and performance

of the steam turbine is obtained by a different approach. Backed by experimental

data, Table 1, and rules of current practice, [59], it is well known that in an off-design

operation of a steam turbine the volumetric flowrate profile of the steam has to be

identical to the nominal volumetric flowrate profile. The reasons and implementation

of this approach are detailed next.

Experimental data from a standard Rankine cycle without carbon capture and

sequestration (CCS), with power ratings similar to those of the pressurized OCC

process, are shown in Table 4.1. The specifications of the working fluid and the

response of each of the turbines are shown for different thermal loading. The presented

operations are not necessarily optimal, but the efficiency of the expansion line for the

particular values of the working fluid conditions at the inlet of each turbine can be

obtained from the table.

In general, the exergy of the thermal energy transferred from the flue gas to the

working fluid increases with increasing temperature of the working fluid. Therefore,

it is favorable to maintain the main stream and the reheat stream temperature at

the highest possible value, 600'C and 610'C respectively; moreover, this implies that

the inlet temperature of the turbines are at their nominal values. Now in order to

maintain the temperature of the working fluid at the exit of the HRSG/inlet of the

HPT and IPT constant despite the decrease in the thermal load, the mass flowrate

of the working fluid has to decrease as seen in Table 4.1.

Higher pressures of the working fluid increase the exergy of the process, however,

the characteristics of the turbine expansion line necessitates decreasing the working

141
fluid pressure with the decrease in thermal load. More specifically, turbines must

maintain an approximately constant flow pattern for an efficient and reliable opera-

tion, [59], where steam flows smoothly over the blades' surfaces rather than colliding

with them. The turbine angular velocities and thus the blades velocities are constant

besides during startup and shutdown. Therefore, for fixed blades design, to achieve

the constant design flow pattern, a constant volumetric flowrate of the working fluid

through any section of the turbine is required; note that the volumetric flowrate of the

working fluid increases as the steam expands through the turbine. Variations in the

volumetric flowrate profile result in deterioration of the lifetime and efficiency of the

turbines. To achieve constant volumertic flowrate profile at reduced mass flowrate,

the inlet pressures of the turbines are decreased. This can be achieved by (i) a con-

stant pressure at the steam generation followed by throttling at the turbine inlet or

(ii) so-called sliding pressure boiler, i.e., variable pressure at the steam generation,

[60, 61]. Herein, it is assumed that the HRSG generates steam at sliding pressure.

In the data of Table 4.1, the volumetric flowrate of the working fluid at both the

inlet and exit of the turbines is constant for different loads except for exit of the low

pressure turbine which is maintained at the nominal condenser's pressure. To achieve

this in modeling, the inlet and outlet pressures of the high pressure and intermediate

pressure turbines are taken as function of the working fluid mass flowrate. In other

words, the equation V = rh x v (P,T) = constant is enforced at the inlet and outlet of

each turbine section, where V and v are the total volumetric flowrate and the specific

volume, respectively, while rh, T, &P are the mass flowrate, temperature, and pres-

sure of the working fluid. The temperature is fixed at inlet, as explained above, and

is a dependent variable throughout the expansion process. Thus, the pressure and

the mass flowrate are no longer independent resulting in a single degree of freedom

in the above equation. Herein, the mass flowrate is chosen as the degree of freedom,

for numerical reasons. The mass flowrate is an optimization variable and its value

142
Table 4.1: Operating conditions of working fluid and turbine expansion line at differ-
ent loads. Each turbine operates at a constant volumetric flowrate profile (constant
volumetric flowrate at each section of the expansion line). The outlet pressure of
the LPT, condenser pressure, is constant and equal to 0.042bar for the assumed wet
cooling condenser

7
Thermal Load % Pinlet bar inlet temperature 'C 7nFW, actual t
lactual Pinlet
FW, nominal lnominal Poutlet

High Pressure Turbine (HPT)


100 250 600 100 100 4.2888
70 169.9 600 64.7 99.44 4.177
50 120.1 600 45.6 99.33 4.093
35 84.62 600 32.0 98.20 4.058
Reheat - Intermediate Pressure Turbine(IPT)
100 52.90 610 100 100 5.204
70 36.91 610 69.8 98.75 4.98
50 26.55 610 50.3 98.75 4.80
35 18.95 610 35.8 98.75 4.68
Low Pressure Turbine(LPT)
100 10.44 - 100 100 239.5
70 7.60 - 72.9 99.25 174.4
50 5.67 - 53.9 98.18 130.2
35 4.15 - 39.3 96.79 95.3

is determined by energy balance during optimization of each thermal load. For each

turbine section, the efficiency ratio, the inlet pressure ratio, and the outlet pressure

ratio are mapped versus the mass flowrate ratio, wherein each ratio is taken relative

to the nominal. With very high accuracy, the inlet and outlet pressure ratios are

affine linear functions of the mass flow ratio, as a direct consequence of the physical

properties of the water particularly in the very superheated state after the reheat

where temperature is relatively high and pressure is relatively low. The efficiency ra-

tio is considered as a piecewise linear function versus the working fluid mass flowrate

ratio because it does not seem to follow any particular continuous function versus the

mass flowrate ratio.

143
4.4 Modeling Approach

4.4.1 Process Operating Parameters

Similar to [24, 27, 38, 35] and Chapters 2&3, the model is implemented in Aspen-

Plus. Figure 4-1 represents the process where a surface heat exchanger (RHE) is

used for sensible and latent heat recovery from the water in the flue gas. Note that

the analysis is pseudo-steady state, i.e., does not consider the transition dynamics

between different power levels. The variables and constraints on Figure 4-1 are used

for optimization.

For the scope of the current work, the air separation unit (ASU) is accounted

for by its power requirements and is not modeled rigorously. Operating the ASU at

partload increases the specific power requirement for producing the oxygen stream.

However, herein the power demand and specifications of the ASU are considered

constant (for any design and operation) for two reasons: first, a design specific to a

certain load is required to operate within the complete range of operation of the power

plant (30%-100%), thus the ASU should be capable of supplying the full load oxygen

requirements; therefore, the ASU design is the nominal/full-load design. Second, it

is assumed that oxygen storage is possible; therefore, for any dedicated design the

specific power demand of the ASU is identical to that of the nominal operation and

design. It is worth mentioning that even if storage is not practical then the ASU is

still required to have the full load design, and therefore using an elaborate model for

the ASU leads to changing the results in terms of values but not in the terms of the

trend or relative behavior. In particular, the comparison of the flexible design to the

dedicated designs will not be affected, which is the aim of this study. In conclusion,

the power demand and specifications of the ASU are considered constant for any

design and operation. In other words, the effects of load variation on the ASU are

not accounted for without jeopardizing the conclusions of flexibility to the thermal

144
load.

Pressurized ASU
02 02
o pressor Air
IP I 4-
L 02+Ar separat
N2

-R c-prn FW-HRSG-1n
FG-Rec-sec

om-Gas-in "ReheatRH

mbustionre-Recovery Unit dSequestrated C02


Coa Wat
Slurry Gas ot-G o aG-e vou

Contrller C ITvented gas


Ah A
FW-Recov-in

Condensate

P~ ~~ ra atio.

IBL DI .Flw leed Stage3 ., Bleed IBLD3 Flowl


-B Bleed-

Reheatq rtor Cooling Water

FW-HRSG-in FW FW , Condenser

HP-Pump

LP-Pump

Figure 4-1: Oxycombustion cycle flowsheet based on wet recycling. Note that this
schematic does not represent entirely the modeling, e.g., turbines were modeled with
multiple sections in AspenPlus®

Table 2.1 shows the fixed parameters of the process, pertaining to the operation

of the Isotherm® combustor, [30], and other components. To satisfy these while

varying the load requires adjustments to several streams in order to satisfy the ope

ration requirements. First of all, changing the thermal load requires a change in the

amount of coal and thus the amount of slurry water needed to transport the coal

keeping the required water ratio in the coal water slurry mixture. Moreover, the

amount of oxygen required to oxidize the fuel also changes, and therefore, the air

flowrate entering the air separation unit changes. Also, the amount of the atomizer

145
stream needed to atomize the coal water slurry entering the combustor changes. In

the Rankine cycle, the steam leaks from the turbines are assumed to scale linearly

with the working fluid flowrate. Now that the atomizer stream flowrate, which is

extracted from the steam expansion line, and the turbine leaks change, the amount of

makeup water also changes. All of the dependent variations are modeled in calculator

blocks or design specifications in order to maintain the operation constraints of the

process at any load.

4.4.2 Flue Gas Pressure Losses

Another important aspect of the model is the calculation of the flue gas pressure

losses as it passes through the HRSG and the recycling pipes. Pressure losses are

calculated by utilizing standard pressure loss equations [62, 631, and similarity analy-

sis [24, 27, 35, 38] and Chapters 2&3. The losses depend on the designs of the HRSG

and the recycling pipes, which depend on the process design pressure and economical

considerations [24, 271. Therefore, at a given thermal load, the pressure drops depend

on whether the HRSG and pipes are designed for that specific load, or are operated at

a load different than their design load. In the case of the former, the process pressure

and the HRSG and pipe designs are allowed to change upon design (PComb, design is

a design variables, Section 4.5.1), and the representation of the pressure losses are

identical to those of [24, 271:

Design specific for a particular load

For the recycling pipes upon design:

L V2
AP,3e=p
=

146
where V is the bulk gas velocity in the pipe, d is the pipe diameter, LP is the pipe

equivalent length, p is the gas density, and f is the friction factor calculated by

- -2
2 E/d)
L(2log)(2e/d) 2o)J13
= 'p-2. log 7.4 Red 7.4 Red

where E is the pipe roughness, Red = PVd is the Reynolds number based on the pipe

diameter, p is the flue gas density, and p is the dynamic viscosity of the gas.

The pipe diameter, d, and the gas velocity, V, are related by mh = pV-2-, where ni is
the recycled gas mass flowrate through each of the two pipes. Note that for practical

considerations the pipes diameters and the gas velocities in the pipes have to fall

within fixed ranges. The acceptable ranges are shown in Table 4.2 similar to those

considered [21, 24, 27]. The equivalent length is obtained by considering a 63.5 mbar

pressure drop in each pipe at an operating pressure of 10 bar based on experimental

data from ENEL.

Table 4.2: Recycling pipes diameters and gas velocity ranges, [21, 24]

Diameter Range (m) Velocity Range (m/s)


Primary Recycling Pipe [1-6] [4-25]
Secondary Recycling Pipe [1-4.15] [14-30]

The larger the pipe diameter, the smaller the gas flow velocity, and the smaller

the pressure drop. Thus, a larger pipe is always favored in terms of efficiency but not

necessarily from an economical point of view as the capital, installation, and mainte-

nance costs would increase. However, very large flow velocities can cause structural

failure and acoustic resonance. Herein, at each iteration within the optimization

study, the largest allowable diameter, for each of the two pipes, is chosen such that

the gas velocity remains within the velocity range. In the case that the flowrate is

too high, the upper bound on the diameter is chosen and the velocity range is violated.

147
For the HRSG upon design:

APHRSG,a _HRSG,a a hb
APHRSG,b QHRSG,bPb a

where QHRSG is the rate of thermal energy transferred in the HRSG, rh is the flue gas

mass flowrate, and a and b stand for actual design and basecase design respectively.

Operation with a fixed design

On the other hand, when the process operates at a load while designed for a different

load, the specifications of the HRSG and the recycling pipes are identical to those

of the design load. Moreover, the design pressure is the maximum allowed pressure

of any operation. The losses are computed based on similarity analysis between the

actual, a, and design, d, operating conditions:

The recycling pipes

y (f Lp &,
p
APpipe, a _( d' 22 )a ( 7 h2)a _ d aPd

A pipe, d (d3d ( 2)d fadPa

The pipes lengths, Lp, and the pipes diameters, d, are constant and equal to those

of the design load, therefore, do not appear in the final expression.

For the HRSG:

APHRSG, a (fNp2)a y__2_a


APHRSG, d (fNp vj,) d (fPV6) d

m 2 p
fp 2A d

dPa

148
The HRSG design and size are invariant from those of the design load, therefore, the

number of rows N, and the cross sectional area Ac, are constant. Moreover, the tube

diameters D, the transverse and longitudinal distance between the tubes ST&SL, are

also constant leading to _ = . f is the friction factor, which is approximately

constant for the ranges of Reynolds numbers involved. The pressure losses at a given

operation load are inversely proportional to the density of the flue gas. In other

words, optimization of operation while the process is designed for a different load

results in the maximum allowable value of the operating pressure, since it decreases

pressure losses and does not decrease thermal recovery; also the tradeoff between the

02 and the CO 2 compression is insensitive within those ranges of pressures. However,

since the operation pressure cannot exceed the design pressure of the design load, the

optimum operating pressure is expected to be equal to the design pressure; this is

verified by the results.

4.5 Optimization Formulation

The objective of the study is to achieve a high performance for the pressurized OCC

process subject to variable loading. In particular, it is first desired to determine the

flexibility of the process. If the performance of the process designed for the nominal

load but operating at a different load is lower than the performance of the process

designed specifically for the non-nominal load, then the process is not ideally flexible.

In that case, the original design needs to be revisited, such that a maximum overall

performance is achieved rather than maximum performance at nominal conditions. To

examine the ideal flexibility, two families of optimization processes are required over

the range of possible thermal load; first, optimization of operation under the nominal

design, and second, simultaneous optimization of design and operation at the specific

thermal load. Similar to the study of uncertainties at nominal loading, [351 and

149
Chapter 3, a classification of the optimization variables as design and as operation

is required. Operation variables can change with load variations, whereas, design

variables are fixed. As described in Section 4.5.2, load variations impose additional

constraints compared to nominal operation.

As mentioned above, the redesign of the process for a specific load is considered to

allow for the same range of load operation; therefore, the turbines are ones that can

accommodate the nominal load operation, without reaching unrealistic working fluid

pressures, thus have the nominal load design. Recall that an expansion line design

requires maintaining the design flow pattern for maintaining high turbine performance

and reliability. More specifically, the inlet pressures and pressure ratios of the turbines

are adjusted as a function of the working fluid flowrate in order to maintain the

fixed volumetric flowrate profile of the working fluid in the turbines. Operating at a

load lower than the design is possible by decreasing the turbines' inlet pressures and

pressure ratios, and vice versa. Conversely, to operate at higher than design load, the

pressure would have to be increased. But to achieve high performance at nominal

conditions, the pressure is typically set at the maximal allowed limit, thus further

increase is not possible. So if the expansion line is designed for a partload operation,

then the plant is incapable of providing the initial full load power rating. However,

since a power plant is expected to operate for the majority of its lifetime under the

initial full load power rating, then the full load design of the expansion line is the

desired design for any partload.

4.5.1 Design and Optimization Variables

Table 4.3 characterizes the variables as design or operation and provides their range,

and Figure 4-1 marks the variables on the flowsheet. The basecase default values are

the optimized results of the design that is ideally flexible to uncertainties in coal, am-

bient conditions, and input stream specifications of the process at nominal load [35].

150
The classification of the variables is similar to that in [35] and Chapter 3 with few ad-

ditions and differences. The methodology of implementing active constrains depends

on the variables' characterizations and is discussed in Section 4.5.3.

Combustor Pressure, PComb, o & PComb, d

As discussed in Section 4.4.2, the conbustor's operating pressure, PComb, o, is an op-

eration variable, while the upper bound on its range, equal to the design pressure of

the design load, PComb, d, is a design variable. It is also seen that for a fixed design,

the pressure losses are smaller with increasing operation pressure, and therefore, op-

timization is expected to choose the maximum allowed value i.e., the design pressure.

This is verified by the results.

Combustor's Duty, QComb

In [35] where only the nominal load is considered, the combustor's duty, QComb, is

treated as a design variable. However, the combustor's specifications and insulation

are fixed with the design, and the value of QComb for a given design changes with the

operating load. At a given design with a QComb, d, the resulting actual operational

combustor duty, QComb, a, is assumed to scale with the design duty according to

the ratio of the thermal loads of the design, d, and operation, o; in other words,

Comb,. Note that the linear dependence is consistent with the extreme

cases: for a fixed design, at zero load there is zero transferred duty, and at design

load the transferred duty equals to the value at design. Similarly, the range of QComb, d

is assumed to depend on the nominal duty QComb, n according the following equation,
QComb, d _ TLoadd _ TLoadd
QComb, n TLoadn 1

Feedwater Flowrate, rnFW, Main

The performance curves of the turbines are available as discussed in Section 4.4. The

pressures of the working fluid are set to accommodate the turbine expansion line con-

151
straints. The main feedwater flowrate is an operation variable with no restrictions on

the range of its variation. It is trivial to deduce that the optimal flowrate decreases

with decreasing load. Therefore, bounding the range of the feedwater flowrate from

above at the nominal value facilitates the problem without constraining the optimiza-

tion; a slightly larger upper bound is used nevertheless to make sure that the problem

is not constrained.

Bleeds' Flowrates, rnBLD1,2,&3

The bleeds' flowrates can, and should, change with different thermal loads since the

amount of regeneration required for the varying feedwater flowrate changes. More-

over, the effectiveness of regeneration changes due to the change in the expansion line

pressures, as explained later, and require a different optimum flowrate. Therefore,

rnBLD1,2,&3 are considered as operation variables.

Bleeds' Pressures, PBLD1,2,&3

In [35], where only the nominal load is considered, the bleeds' extraction pressures

are fixed with the initial design of the turbine expansion line, representing a fixed

position of extraction within a turbine stage, and thus are design variables. Here

however, fixed bleed extraction position is not equivalent to fixed extraction pressure

since at different loads the pressure ranges of each turbine stage change. Therefore,

a fixed extraction position is the required design variable, and this results in different

bleed pressures at different loads. Recall that for a given design, the turbines operate

with a fixed volumetric flowrate throughout in order to maintain the efficient flow

pattern. Therefore, a fixed extraction position is one that has a fixed volumetric

flowrate of the working fluid passing through the turbine at the section of extraction.

Modeling the bleed extraction positions by specifying the bleed pressure that

satisfy the volumetric flowrate constraint would result in a highly iterative procedure

and in an extremely complicated optimization process.

152
Fortunately, it is possible to eliminate the need for iteration and reproduce the

fixed extraction position by simple formulas. To achieve this, the above iterative

procedure is separately implemented on an isolated expansion line, and the results

are compared to fixing the ratio of the extraction pressure relative to the inlet and the

outlet turbine pressures. The comparison is performed over a range of the working

fluid flowrate, representing a range of operation load. At the range of interest, keeping

the simple ratio of bleed extraction pressure, namely Pex"r"ctionPo"t"et as constant over

a range of 100% to 30% load, results in less than a 0.5% difference in the volumetric

flowrate of the working fluid at the point of extraction. This implies that keeping the

pressure ratio fixed according to the dimentionless variable proposed, corresponds to

a fixed extraction position. Keeping the pressure ratio fixed is easy to implement and

makes optimization tractable.

Extraction Stage, Integer Variables

Similar to [24, 38, 351, integer variables are used to select the extraction stage, where

each stage has different ranges of operation and different performance properties. The

discrete variables representing the extraction stage are also design variables.

FWH Duty Transfer, QFWny&2

The duty transfer within each FWH is an operation variable.

Deaerator Pressure, PDeaerator

The deaerator operating pressure is an operation variable determined primarily by

the low pressure pump delivery pressure and the deaerator bleed pressure at desti-

nation. Due to load variation the bleeds' pressures are not constant, and thus the

deaerator pressure is subject to change. However, upper and lower bounds on the

deaerator pressure should be respected. A deaerator operating temperature range

of 101 to 200'C is common practice [64]. This temperature range is equivalent to a

153
pressure range of 1.076 to 15.55 bar. In other words, the pressure should be higher

than atmospheric (1.013bar), and lower than 15.55bar, the pressure of the maximum

allowed saturation temperature of 200*C, which is a material constraint. Regarding

the low-pressure feedwater pump, the pressure range of 5-15 bar obtained in the re-

sults, Section4.6, is considered within the acceptable range of operation. To achieve

variable pressure, a variable-speed driven (VSD) pump could be utilized. If instead a

fixed-pressure pump is utilized, then the pressure of the low-pressure pump can be set

to around 15 bar (maximum pressure required for any operating load, Section 4.6) and

the feedwater would be throttled into the deaerator tank (if and when needed). The

power requirement increase due to always operating at 15bar is minor, less than 0.01

percentage points change in efficiency of the cycle, because the pumping requirements

are relatively low.

Temperature of Flue Gas Exiting the RHE, TFG-RHE-out

The temperature of the flue gas exiting the RHE, TFG-RHE-OUt, is also an operation

variable dictating the amount of recovered thermal energy. However, the optimal

value is not expected to vary with different operating loads because the optimum flue

gas operating pressure is high enough to allow the pinch point to occur at the flue gas

exit from the RHE. This expectation is confirmed by the obtained results. Note that

even if the value of TFG-RHEOut is invariant, the amount of recovered thermal energy

at the RHE, QRHE, changes because different loads have different flue gas flowrates

and different amounts of energy to recover.

Additional Integer Variables Required for Partload Optimization

Finally the process has three additional binary variables, each denoting the activation

or deactivation of a bleed flow. In principle, deactivating the bleed flow is identical to

setting the bleed flow to zero, however, the binary variables are required for modeling

purposes: zero flowrates cause convergence and mass balance errors in AspenPlus.

154
Table 4.3: Design and operation variables. The integer variables BLDAflow
inhibit or allow a bleed flow; inhibiting a flow by a BLD-flow value of zero is,
in practice, equivalent to setting 7hBLD to zero, but required here for modeling
purposes

Number Variable Range Variable Type


1 PComb, operation [1 - PComb, design] bar Operation
2 PComb, design [1 - 10] bar Design
3 (c)omb [1 - 30] x TLOAD MW Design
4 mFW, Main [20-300] kg/s Operation
5 BLD1_stage integer Stages: 1-4 Design
variable
6 PBLD1--BLD.stage, outlet
PBLD1..-tage, inlet -PBLDL.stage, outlet
[0 - 1] Design
7 ThBLD1 [0 - 60] kg/s Operation
8 BLD2..stage integer Stages: 3-6 Design
variable
9 DM2-PBLD2-stage, outlet [0 - 1] Design
PBLD2.stage, inlet -PBLD2..stage, outlet
10 rnBLD2 [0 - 30] kg/s Operation
11 BLD3_stage integer Stages: 5-7 Design
variable
12 PBLD2-PBLD3.stage, outlet
BLD3..stage, inlet -PBLD3.stage, outlet
[0 - 1] Design
13 mBLD3 [0 - 30] kg/s Operation
14 QFWH1 [0 - 200] MW Operation
15 QFWH2 [0 - 200] MW Operation
16 PDeaerator [2 - 15.55] bar Operation
17 TFG-RHEout [30 - 200] 0C Operation
Required for partload optimization
18 BLD1-flow binary {0,1} Operation
variable
19 BLD2-flow binary {0,1} Operation
variable
20 BLD3-flow binary {0,1} Operation
variable

The three integer variables are BLDiflow, one for each bleed, and are operation

variables, i.e., can be activated depending on the load. Physically these variables are

controlled in the same manner the bleed flow is controlled.

155
4.5.2 Constraints

Constraints on the admissible design and operation are imposed based on physical,

practical, and economical considerations. The turbines' volumetric flowrate profile

is a major constraint which is satisfied by automating the turbines' inlet pressures

and pressure ratios. Twelve constraints are accounted for explicitly at any operation

load. The constraints are listed in Table 4.4 and illustrated in Figure 4-1. Further,

active constraints presented in Section 4.5.3 allow satisfying the constraints at the

simulation level, while providing optimal performance.

Table 4.4: Optimization constraints, [24, 27, 35] and Chapter 3. The
most challenging constraints are eliminated by setting the bounds on
the variables and automating the turbine inlet pressures, turbines pres-
sure ranges, and bleeds extraction pressures, and thus not shown here

Nu mber Constraint Value


1 MITAHRSG > 3.7 0C
2 AreaFWH1 7,069m 2 equivalent to
0
MITAFWH1=2.1 C at
nominal
3 AreaFVWH2 5,254m 2 equivalent to
MITAFWH2=2.1'C at
nominal
4 qDeaerator Saturated Liquid
5 PDeaerator * 2bar
6 PDeaerator ; 15.55bar
7 TCool-Gas 20'C above acid condensa-
tion temperature
8 TFW-HRSG-in 50 C above acid condensa-
tion temperature
9 TCom-Gas-in 20'C above acid condensa-
tion temperature
10 MITARHE > 7-5 0 C
11 CO 2 -Cap 94% of total CO 2 produced
required to be captured
12 CO 2 _Pur captured CO 2 is 96.5% pure

156
4.5.3 Active Constraint Optimization

In [24, 27, 28, 29] and Chapter 1 it is proven that optimal operating conditions occur

at some active constraints. Enforcing these constraints as operation specifications,

facilitates the optimization in several aspects: (i) avoid constraints violations, (ii)

avoid simulation errors and failures, (iii) accelerate convergence, (iv) avoid conver-

gence to suboptimal local optima. The desired active constraints can be satisfied at

the simulation level by manipulating the main influencing variables.

The following constraints and variables are coupled:

1. MITAHRSG /ThFW, Main: The allowed minimum internal temperature approach

constraint on the HRSG is achieved by manipulating the main feedwater flowrate

2. MITARHE/TF-HE-out: The allowed minimum internal temperature approach

on the RHE achieved by manipulating the temperature of the flue gas exiting

the RHE

3. & 4. Double-pinch.FWH(1&2)/FWH(1& 2) and rnBLD(1&2). Both the duty trans-

fer within each closed FWH and the flowrate of the respective bleeds are utilized

in order to guarantee equal values of the temperature approach at the feedwater

heater outlet, and at the position of phase change of the bleed, [28, 29] and

Chapter 1. Note that the areas of the feedwater heaters are fixed and thus the

value of the double-pinch is not necessarily equal at different loads:

5. PDeaerator: For optimal operation, the deaerator pressure has to be equal to

the pressure of the deaerator bleed, BLD3, at the deaerator inlet; [28, 29] and

Chapter 1. Therefore, PBLD3 and PDeaerator are coupled to be equal at the

level of the deaerator, i.e., after accounting for friction and hydrostatic pressure

changes. The deaerator bleed flowrate also plays a role in the optimum value

of the deaerator pressure since it affects the amount of pressure loss in the

connection pipes

157
6. qDeaerator/nBLD3- The mixture inside the tank deaerator has to reach the satu-

rated liquid state for the effective removal of dissolved air in the working fluid.

This constraint is satisfied by the low pressure bleed to the deaerator

Note that the last two items are relevant only when the deaerator bleed is active,

i.e., BLDJflow3=1; otherwise when BLD.flow3=0, the constraint on the quality of

the mixture inside the dearator is satisfied by manipulating the deaerator pressure.

Moreover, the combustor operation pressure is not equated to the design pressure in

an active constraint, in order to show that indeed the optimum operation pressure is

the process's design pressure.

4.6 Results and Analysis

Three steps are performed to examine the partload performance of the pressurized OC

process. First, a given partload is operated with the nominal design and operating

conditions, [35] and Chapter 3, with changes in the values of the deaerator pressure,

feedwater flowrate, and FWHs' duty, in order to satisfy the constraints set on the

deaerator pressure and satisfy the FWHs areas. Then, optimization of the operation

is performed while keeping the same fixed design. Third, the process is redesigned for

that specific partload by simultaneous optimization of design and operation. Recall

that the turbine expansion line design is kept fixed in this step, but not the design

of the extraction bleeds' positions. The latter optimization obtains the maximum

possible performance at that specific load (for the given turbines). Comparing the

results of steps two and three, the flexibility of the process to varying load is obtained.

The results show that the process is ideally flexible to load variations. The flexibility

eliminates the need for a prior knowledge of the operation schedule. The reasons

behind the ideal flexibility are shown in Sections 4.6.2, 4.6.3 & 4.6.4. The comparison

of steps 1 and 2 shows that optimization of operation is required even for a flexible

158
design, to achieve high performance. Comparing Steps 2 and 3 show that a flexible

design can match the optimum performance of the process at the evaluated variable

loads.

In the case study considered, the lowest thermal load is 35%, see Table 4.1. Herein,

flexibility assessment is performed om thermal loads reaching down to 30%, to account

for extreme scenarios. Table 4.5 summarizes the results of the RHE configuration

of [351 and Chapter 3, which is the ideal flexible design to uncertainties of coal,

ambient conditions, and input streams specifications at nominal load, and serves

as the basecase of this study. Tables 4.6&4.7 present the results of the flexibility

evaluations for the 60% and 30% thermal loads, respectively.

4.6.1 Flexibility Assessment

The results of the flexibility assessment, Tables 4.6 & 4.7, show that the process

designed for full load is ideally flexible to large load changes. By optimizing the

operation of a design specific for the full load while under a partial thermal load

(columns 3 in Tables 4.6 & 4.7), the process performance matches the maximum
possible performance obtained when the process is designed specifically for that load

-apart from the turbines- (columns 4 in Tables 4.6 & 4.7). The discrepancy is ap-

proximately 0.02% and as such insignificant. In contrast, maintaining the nominal

load operating conditions when at a partload operation (simulations in columns 2 in

Tables 4.6 & 4.7), results in a drastic decrease in the efficiency of the process, around

8 percentage points decrease at the 60% TLoad, and 30 percentage points 30% load.

Figure 4-2 represents the results of all the operation ranges graphically, where the

optimum operation of the full load design at a particular load matches the respec-

tive optimal design of that load, while without optimization the performance suffers

significantly and may even be infeasible. Note that the 90% load has an efficiency of

159
Table 4.5: Summary of optimal design and op-
eration of the RHE process for coal and ambi-
ent conditions variations and FWH area spec-
ification, [35] and Chapter 3 Table 3.10. Re-
sults shown are for the coal considered herein

Fuel Flowrate 30kg/s


Slurry water flowrate 16.50kg/s
Atomizer Stream 2.50kg/s
flowrate
Air flowrate 311.2kg/s
Efficiency (based on 34.41%
LHV)
Independent and Key Dependent /araibles
PComb 7.41bar
1MW
rnFW Main 306kg/s
PBLD1 99.Obar
mBLD1 62.2kg/s
PBLD2 26.Obar
ThBLD2 14.7kg/s
PBLD3 16.8bar
ThBLD3 9.27kg/s
PDeaerator 14.7bar
OFWH1 138MW
QFWH2 37.7MW
TFG-RHE-out 36.9 0 C
RHE 122MW
Depend ent Variables
Tcool-Gas 327 0 C
TFW-HRSG-in 3220 C
TComb-Gas-in 3090 C
Flue gas fkiwrate in 1,138kg/s
HRSG
Condensed water 33.2kg/s
RHE
APHRSG 0.265bar
APp, Primary 0.054bar
APp, secondar 0.035bar
rnFG-Rec-pri 281kg/s
rnFG-Rec-sec 737kg/s
Rankine cycle net 409MW
power
TFW-Recov-out 1640C
PHPT 250bar
PReheat 53.5bar
Table 4.6: 60% thermal load flexibility

Variable 60% thermal load 60% thermal load 60% thermal


simulation optimization of load optimization
operation with of design and
nominal load operation
design
Input Parameters
Fuel Flowrate 18kg/s 18kg/s 18kg/s
Slurry water flowrate 9.9kg/s 9.9kg/s 9.9kg/s
Atomizer Stream flowrate 1.5kg/s 1.5kg/s 1.5kg/s
Air flowrate 191.7kg/s 191.7kg/s 191.7kg/s
Efficiency (based on LHV) 23.70% 31.59% 31.61%
Independent and Key Dependent Varaibles
PComb 7.41bar 7.41bar 7.41bar
OComb 0.6MW 0.6MW 0.6MW
mFW Main 179kg/s 168kg/s 169kg/s
PBLD1 64. Ibar 60.4bar 64.6bar
equivalent PBLD1 at nominal 99.Obar 99.Obar 105bar
rhBLD1 62.2kg/s 27.9kg/s 29.3kg/s
PBLD2 16.3bar 17.Obar 18.8bar
equivalent PBLD2 at nominal 26.Obar 26.Obar 28.0bar
ThBLD2 14.7kg/s 7.27kg/s 8.60kg/s
PBLD3 8.17bar 11.0bar 11.2bar
equivalent PBLD3 at nominal 16.8bar 16.8bar 17.0bar
rhBLD3 9.27kg/s 4.33kg/s 4.10kg/s
PDeaerator 8.16bar 10.3bar 10.6bar
FWHV1 87.0MW 67.2MW 67.9MW
QFWH2 27-4MW 19.0MW 22.7MW
TFCRHE-out 93.8 0 C 36-90 C 36-9 0 C
RHE 58.2MW 70.1MW 70.1MW
Dependent Variables
TCool-Gas 2 311 0C 2960 C 301 0 C
TFW-HRSG-in
0
306 C 291 0C 296 0 C
TComb-Gas-in 296 0 C 283 0 C 288 0 C
Flue gas fkiwrate in HRSG 665kg/s 648kg/s 637kg/s
Condensed water RHE 17.1kg/s 19.9kg/s 19.9kg/s
APHRSG 0.334bar 0.334bar 0.334bar
APpipe Primary 0.035bar 0.034bar 0.03bar
APpi,, secondary 0.009bar 0.008bar 0.008bar
mFG-Rec-pri 167kg/s 165kg/s 166kg/s
FG-Rec-sec 426kg/s 411kg/s 416kg/s
Rankine cycle net power 196MW 242MW 239MW
TFW-PRecov-out 173 0C 157 0 C 158 0 C
PHPT 161bar 151bar 152bar
PReheat 42.6bar 39.7bar 40.Obar
Table 4.7: 30% thermal load flexibility

Variable 30% thermal load 30% thermal load 30% thermal


simulation optimization of load optimization
operation with of design and
nominal load operation
design
Input Parameters
Fuel Flowrate 9kg/s 9kg/s 9kg/s
Slurry water flowrate 4.95kg/s 4.95kg/s 4.95kg/s
Atomizer Stream flowrate 0.75kg/s 0.75kg/s 0.75kg/s
Air flowrate 102kg/s 102kg/s 102kg/s
Efficiency (based on LHV) Infeasible 24.48% 24.48%
Independent and Key Dependent Varaibles
PComb 7.41bar 7.41bar 7.41bar
QComb 0.3MW 0.3MW 0.3MW
mFW Main 74.5kg/s 74.5kg/s
PBLD1 31.8bar
equivalent PBLD1 at nominal 99.Obar 99.Obar 106bar
ThBLD1 9.41kg/s 8.75kg/s
PBLD2 8.66bar 10.3bar
equivalent PBLD2 at nominal 26.Obar 26.Obar 29.9bar
ThBLD2 3.23kg/s 4.19kg/s
PBLD3 NA NA
equivalent PBLD3 at nominal 16.8bar 16.8bar NA
rhBLD3 0kg/s 0kg/s
PDeaerator 5.04bar 5.23bar
QFWH1 24.6MW 22.7MW
QFWH2 8.93MW 11.8MW
TFGRiHE-out 36-9 0 C 369 0 C
RHE 33.1MW 33.3MW
Dependent Variables
TCool-Gas 253 0 C 256 0 C
TFW-HRSG-i 247 0 C 250 0 C
TComt-Gas-in 246 0C 248 0C
Flue gas fkiwrate in HRSG 302kg/s 304kg/s
Condensed water RHE 9.96kg/s 9.96kg/s
APHRSG 0.334bar 0.334bar
APi, Primary 0.008bar 0.008bar
AP 2, secondary 0.002bar 0.002bar
hFCRec-pri 80.2kg/s 80.3kg/s
FG-Rec-sec 186kg/s 188kg/s
Rankine cycle net power 110MW 109MW
TFW-Recov-out 157 0 C 1500C
PHPT 72.lbar 72.4bar
PReheat 19.8bar 19.9bar
34.55%, larger than the full load efficiency because the benefits of smaller pressure

losses and smaller flue gas flows outweigh the disadvantages of the turbines expansion

and turbines bleeds reduction in performance; this validates that the full load design

takes economic consideration into account for the sizing of the HRSG and recycling

pipes [24, 271.

35
32.5 -
30
S27.5-
25- -
40
22.5
.0 20-
' x
... Simultaneous optimization
4 17.5- I of design and operation
Optimization of operation
15 - for a fixed (full-load) design
12.5- 4- Simulation of the fixed
(full-load) design
10
0.2 0.4 0.6 0.8 1
Thermal Load
Figure 4-2: Simulation and optimization of the operation for the full load optimal
design at different partloads, in dashed and solid lines respectively. Optimization of
the operation of the full load design matches the performance of the optimal design of
the specific part load, in (x)'s. The full load design is ideally flexible but optimization
is required, while simulation of the full load operation suffers significantly and is even
infeasible at relatively small loads. In all designs, the design of the turbine expansion
line is fixed to allow for the complete range of load operations.

163
4.6.2 Behavior of Key Variables

Analyzing the behavior of the variables is crucial to understand the reasons behind

the flexibility of the process with respect to loading, in particular the variables related

to regeneration and the recovery sections. The behavior of key variables is discussed

in this subsection. In the following subsection, Rankine cycles without pressurized

thermal recovery and CCS are discussed to argue that they do not exhibit ideal

flexibility.

Combustor Pressure, PComb, operation & PComb, design

As mentioned above, the operating pressure at optimum is equal to the upper bound

of its range, which is the pressure of the design load. Moreover, the design pressure

does not change from the nominal design pressure value because the same tradeoff

between thermal recovery and pressure losses that exists at full load, [24, 27], also

occurs here; although higher operating pressures result in more recovery, it results

in larger pressure losses. Also, lower operating pressures reduces the pressure losses,

but achieves smaller thermal recovery at the HRSG. First, the thermo-economical

optimum design occurs at the same range as that of the nominal design. The pressure

is sufficient to allow for the condensation of the majority of the water in the flue gas,

by allowing the minimum temperature approach to occur at the outlet of the flue

gas; note that the pinch also occurs at the onset of water condensation of the flue gas

because a lower pressure eliminates the outlet pinch and a higher pressure results in

larger pressure drop of the designed HRSG and recycling pipes with no change in the

outlet pinch.

Combustor Duty, QComb

QComb is favored to be the minimum allowed value as it reduces the irreversibilities

associated with the heat transfer across a large temperature gradient.

164
Feedwater Flworate, rnFW Main

As expected, the optimal mFW Main decreases with decreasing thermal load. The

amount of thermal energy available for transfer from the HRSG to heat the working

fluid upto the nominal working fluid temperatures, 600'C for the main stream, and

610'C for the reheat stream, results in a smaller working fluid flowrate compared to

the full load operation. At any fixed thermal load, a larger main feedwater flowrate

in general signifies a larger power output and a larger efficiency, columns 3 & 4 of

Tables 4.6 & 4.7. However, this is not the case looking at Column 1 of the same

tables. The main feedwater flowrate is large but the working fluid flowrate decreases

significantly through the expansion line due to the suboptimal high bleeds flowrates,

resulting in a significant decrease of efficiency. Notice that the pressure at the inlet

to the high pressure turbine and the intermediate pressure turbine, PHPT & PReheat

respectively, decrease with the decrease in working fluid flowrate in order to satisfy

the turbines constraints of constant volumetric flowrates. At 60% and 30% thermal

load, the process is subcritical, i.e., there is a phase change of the feedwater in the

HRSG, see Section 4.6.4. It is important to realize that in the two optimized runs of

each load, optimization of operation and optimization of design and operation, the

.npW Main, pload is almost equal to the thermal load;


ratio of the working fluid flowrate, mFW Main, nominal

this fact is due to the thermal recovery section, and is fundamental in explaining the

reasons behind the flexibility. The argument that the recovery section is responsible

for the working fluid ratio relative to nominal being proportional to the thermal load

is proven in Section 4.6.3, but considered as an observation/result in what follows.

Recovery Section (condensed water, QRHE, TFG-RHE.out, TFW-R.H-out)

Before discussing the behavior of the regeneration bleeds, it is important to un-

derstand the thermal recovery section (after the condenser, before the deaerator),

which the working fluid passes through prior to regeneration. The overall flue gas

165
flowrate scales according to the thermal load and has almost identical composition

with the flue gas at nominal conditions; the fuel flowrate, the oxygen flowrate, and

the added water flowrate scale from their nominal values according to the opera-

tion thermal load. Similarly, the overall thermal energy available for transfer also

scales proportionally with the thermal load. Now, due to the invariant flue gas

composition in particular water fraction, the ratio of the low grade thermal energy,

transferred at the RHE/recovery, to the total thermal energy transferred into the

Rankine cycle is almost constant (and equal to the same ratio at nominal load);

note that in the pressurized OCC process the temperature of the flue gas exiting

the HRSG/entering the RHE, Tcbo-Gas, is almost equal in the optimum operations

at different loads. As a result, the amount of recovered thermal energy, at any

load, relative to the nominal load is proportional to the thermal load; in summary,
T*t*1,TI.*d - Tload & 9RHE __ Q
QRHETload = Tload. This is validated
QTotal,100% QTta = QRHE,100%

by comparing the results of QRHE in the optimized runs, or comparing the amount

of water condensed from the flue gas in the RHE (33.2kg/s, 19.9kg/s & 9.96kg/s at

100%, 60% & 30% thermal loads respectively).

Recall that the working fluid mass flowrate in the optimized partload operations

and optimized partload designs above scales with the thermal load and that the con-

denser temperature is constant at any load. Therefore, the extensive properties of

the thermal recovery section at the optimized partloads is almost identical to the

recovery section at nominal load, which is verified by the results; namely the tem-

perature of the flue gas exiting the RHE, TFCRHEout, and the temperature of the

feedwater exiting the RHE, TFW-ReC-oUt. Therefore, a unit mass of the working fluid

leaving the condenser and before reaching the deaerator always experiences identical

conditions irrespective of the load, the turbine expansion line pressure ranges, or the

regeneration bleeds extraction pressures and flowrates. The nominal load recovery is

suited for a power plant with supercritical pressures and large pressure ranges equal

166
to those of the basecase nominal load operation; yet, at partload the recovery is

operating within smaller pressure ranges. In other words, the scaled version of the

thermal recovery section is more than adequate in heating the working fluid during

partload and takes over relatively larger and larger fractions of the working fluid's

gradual preheating processes with decreasing load; thus compensates for the regener-

ation process which inherently and independently faces diminishing performance as

explained next. In conclusion, the thermal recovery in a pressurized OCC process

results in two advantages: first, the maximum efficiency of partload is only slightly

lower than the maximum efficiency of the nominal load, and second, the process de-

signed for nominal load can attain a performance that matches the maximum possible

performance at any load.

Another way to explain the importance of the recovery section is by realizing

that at nominal loading the recovery, with the help of the dearator bleed, is able to

sustain a deaerator pressure of around 14bar, and thus bleed3 is extracted from the

turbine expansion line withing the same range of pressure. At partload, the recovery

is unaffected, and able to deliver the working fluid with similar conditions as those of

nominal, i.e., sustaining the high deaerator pressure; however, at partload, the same

value of pressure for the dearator bleed resembles earlier extraction positions, signi-

fying that the recovery section is contributing to a larger fraction of the working fluid

preheating, and is taking over the regeneration which faces diminishing effectiveness

with decreasing load.

Recall that initially the recovery section replaces a number of low pressure feedwa-

ter heaters and regeneration bleeds found in standard Rankine cycles without pressur-

ized recovery. Unlike the recovery section, the regeneration section is highly affected

by the turbines' pressure ranges and bleeds' extraction pressures. Therefore, a stan-

dard Rankine cycle without a pressurized recovery section is expected to be inflexible

to variable loading, and has a partload performance, whether optimized operation

167
or optimized design, significantly lower than that of the nominal, as elaborated in

Section 4.6.3.

Regeneration Section, Bleeds pressures and flowrates

With the.decrease in the expansion line inlet pressures and pressure ranges, the pres-

sure of a bleed withdrawn from the fixed extraction position decreases, and so does

its saturation temperature. The decrease of the bleed's saturation temperature limits

the temperature rise of the feedwater in the feedwater heater decreasing the quantity

and quality of regeneration. Therefore, with decreasing thermal load and decreasing

expansion line pressure ranges, a unit mass of regenerative bleed extracted from a

fixed position is relatively less beneficial to the unaltered temperature of the feed-

water entering the regeneration section; in the pressurized OCC considered herein

the regeneration section starts with the deaerator while in non-pressurized OCC pro-

cesses it starts at the exit of the low pressure pump after the condenser. The high

temperature source of thermal energy, the flue gas in the HRSG, is still operating at

approximately the same temperature ranges; therefore in general, a decrease in the

effectiveness of regeneration reduces the temperature of the feedwater entering the

HRSG, and causes a decrease in the exergy of the transferred thermal energy at the

HRSG, reducing the efficiency of the process. The effect of the decrease in recov-

ery effectiveness on the Rankine cycle performance is alleviated when a pressurized

recovery section is present as explained next.

The reduction of the regeneration effectiveness at partload is one of the reasons

why in the simultaneous optimization of design and operation of the pressurized

OCC the extraction pressures of the bleeds increase compared to those obtained by

the nominal extraction positions operating at the same load. This signifies that the

extraction positions shift to earlier sections along the expansion line; the smaller the

load, the earlier the designed extraction position is. The second reason behind the

168
earlier extraction positions of the bleeds is that the recovery section covers a rela-

tively larger preheating portion and therefore higher quality regeneration is required;

(as if additional feedwater heaters are introduced before FWH1&2, thus FWH1&2

move upwards in pressure). For example, for bleedi, the nominal extraction position

of 99.Obar at nominal load results in a 60.4bar at 60% partload. After redesigning

for the 60% load, the extraction pressure is 64.6bar equivalent to 105bar at nominal

conditions. Similarly for the 30% load, the optimal design extraction position occurs

even earlier with an equivalent pressure of 106bar at nominal conditions. However,

due to the behavior of the recovery section, which is independent of the thermal load,

there is not much advantage in redesigning the regeneration and extractions for a

specific partload. In other words, while the regeneration effectiveness of the bleeds

and feedwater heaters decrease with smaller loads, which here necessitates earlier and

earlier extraction positions, the effectiveness of the recovery section, relative to the

smaller working fluid pressure ranges, increases and takes over the regeneration sec-

tion. The independence of the thermal recovery section to the turbine expansion line

pressure ranges is the main reason behind the flexibility of the process. This behavior

is best explained by the response of the deaerator bleed/BLD3, where at partload

there is very little interest in a deaerator bleed. In particular, the 30% load requires

no deaerator bleed. This is a definite illustration of how the recovery section takes

on a larger role with the decrease in the thermal load till eventually eliminating one

bleed altogether. Another reason for flexibility, also in favor of the recovery section,

is that the recovery section replaces the low temperature feedwater heaters where the

bleeds' saturation temperatures, and thus bleeds' quality, are highly sensitive to the

bleeds' pressures and the turbine pressure ranges; while the saturation temperature

of the high pressure bleeds, namely bleedsl&2 of FHW1&2 in the pressurized OCC

process, are relatively less sensitive to the pressure ranges. Note that if the conditions

of the feedwater entering the FWHs are altered significantly, like what would be the

169
case if low pressure feedwater heaters are present instead of the recovery section, then

the high pressure feedwater heaters and bleeds require significant change in design

and result in a significantly different performance.

The behavior of the bleeds' flowrates are justified for similar reasons. As the load

decreases the bleeds' flowrates decrease; first, the working fluid has a lower flowrate

and thus requires smaller regeneration. Second, the recovery section takes a larger

portion of the working fluid preheating and less regeneration is required. Finally,

with the decrease in saturation temperature of the bleeds, due to the decrease in the

pressure ranges of the turbines, the possible temperature rise of the working fluid is

smaller and thus smaller bleeds flowrates are required.

4.6.3 Standard Rankine Cycles Without Pressurized Recov-

ery

In Section 4.6.2 it was argued that in the OCC process considered the thermal recovery

section allows ideal flexibility with respect to thermal load. A key to achieve this was

that the mass flowrate of the working fluid is proportional to the load. Herein, it is

argued that this proportionality is enabled by the thermal recovery section and that

Rankine cycles without pressurized recovery section do not show this ideal flexibility.

In Rankine cycles without pressurized recovery, where the boiler is the only site of

thermal energy transfer from the flue gas, the amount of thermal energy transferred

to the working fluid relative to that at nominal approximately scales with the thermal

load; (as seen later, the temperature of the flue gas at the outlet of the boiler in a

partload operation is generally smaller than that at nominal and thus the propor-

tionality is not exact). However, unlike the pressurized OCC process, the working

fluid flowrate does not scale linearly with the thermal load because the quality of the

transferred thermal energy decreases compared to the nominal operation. The argu-

ment is illustrated by contraposition; assume in a Rankine cycle without pressurized

170
recovery that the flowrate of the working fluid scales with the thermal load. The

decrease in the fluid mass flowrate at partload from its nominal value results in lower

turbines' pressure ranges due to the requirement of constant volumetric flowrate. The

regeneration section, which is the only section responsible for the gradual preheating

of the working fluid, decreases in effectiveness relative to the nominal conditions due

to the decrease in the expansion line and bleeds' pressures (even if regeneration is

redesigned for that specific load). As a result, the temperature of the working fluid

entering the boiler is lower than that of the nominal operation. Therefore, part of

the thermal energy transfer at the boiler is required to compensate for the deficiency

in the working fluid's temperature, and therefore by conservation of energy, the mass

flowrate of the working fluid has to be smaller than that originally assumed. The even

smaller flowrate results in even smaller turbines' pressure ranges, smaller regeneration

effectiveness, and smaller temperature of the working fluid entering the boiler, etc.

As mentioned above, this behavior can be considered as a decrease in the quality

of the transferred thermal energy because the average temperature of the feedwater

in the boiler is smaller due to the lower temperature of the feedwater entering the

boiler/exiting the less effective regeneration. To be more rigorous, note that since the

temperature of the working fluid entering the boiler decreases, the temperature of the

flue gas exiting the boiler can decrease too. However, the increased amount of ther-

mal energy transfer due to the larger flue gas temperature drop is smaller than the

amount of thermal energy required to elevate the smaller temperature of the work-

ing fluid entering the boiler because the thermal capacity of the flue gas is smaller

than that of the working fluid, guaranteeing that the mass flowrate of the working

fluid relative to the nominal scales sublinearly with the thermal load. The fact that

the thermal capacity of the flue gas is smaller than that of the working fluid can be

deduced in several ways, (i) the pinch inside the HRSG occurs at the cold end of the

exchanger (flue gas exit, feedwater inlet), (ii) the temperature drop of the flue gas is

171
much larger than the temperature increase of the feedwater and the reheat streams,

(iii) regeneration increases efficiency, (iv) and of course can be seen by comparing

the temperature profiles of the streams in the HRSG where the hot stream profile is

steeper than that of the cold stream, etc.

Coal-fired Rankine cycles without pressurized recovery have another limitation

for partload operation. At some thermal load the temperature of the working fluid

entering the boiler will fall below the acid condensation temperature, thus further

limiting the heat transfer from the flue gas. This further deteriorates the performance

of the Rankine cycle subject to variable loading.

Moreover, Rankine cycles without pressurized recovery have larger number of

FWHs, particularly low pressure FWHs, compared to the pressurized OCC process

of the same size. The sensitivity of the larger regeneration section to the turbine

pressure ranges, especially at the low pressures, results in a large difference between

the designs of the non pressurized recovery process at different loads. Thus, rendering

the process inflexible and further deteriorating the performance when a fixed design

is operated under variable loading.

In contrast, due to the recovery section in the pressurized OCC process, which

increases in effectiveness and takes over the regeneration section with decreasing load,

the temperature of the working fluid entering the HRSG barely changes, resulting in

the aforementioned linear relation of the working fluid to the thermal load. Nev-

ertheless, the process remains unconstrained by the acid condensation temperature

constraints, and therefore operates at the unconstrained optimum and maintains a

high performance. For further elaboration, using the same argument used for pro-

cesses without pressurized recovery, it can be shown that the optimal ratio of the

flowrate in the pressurized OCC process is actually sustainable, and the results seen

in the optimized runs in the above tables are not coincidental. The increased effec-

tiveness of the recovery section compensates for the decrease in the effectiveness of

172
the recovery section, thus maintaining the temperature of the feedwater entering the

HRSG close to nominal and allows for the scaled ratio of the working fluid.

4.6.4 Partload and Subcritical Operation

As the load and the mass flowrate of the working fluid decrease, the turbine inlet

pressures and pressure ranges decrease to satisfy the turbine expansion line constraints

of constant volumetric flowrate profile. At some flowrate the required working fluid

pressure falls below critical and the process becomes subcritical, where there is a

phase change of the feedwater in the HRSG, and through out that phase change a

constant saturation temperature is maintained. The lower the feedwater pressure,

the lower the saturation temperature, and the larger the enthalpy of vaporization. A

larger enthalpy of vaporization signifies that more thermal energy is transferred to

the relatively low temperature of saturation, decreasing the exergy of the transferred

thermal energy. Moreover, the presence of a larger range where the temperature of the

feedwater is constant results in decreasing the temperature difference at the cold end

of the HRSG and in reaching the minimum allowed temperature difference, which

limits the flow of the feedwater. The efficiency of the 30% load is relatively small

compared to the efficiencies of the 60% load, relatively high saturation temperature,

and to the 100% loads, supercritical, which are not that different from each other.

Note that in non pressurized recovery Rankine processes, the subcritical conditions

are reached earlier/at larger partloads, and the saturation temperatures for a given

subcritical load are lower, than those of the pressurized OCC process; this is due to

the fact that the flowrate of the feedwater decreases at a faster rate in conventional

Rankine cycles compared to the pressurized OCC process.

173
4.7 Conclusion

The flexibility of the ENEL/ITEA pressurized OCC process to variable load is eval-

uated with an accurate representation of unit operations particularly the turbine ex-

pansion line. The turbines operate at constant volumetric fluid flowrate profile which

requires changing the turbine inlet pressures and pressure ranges with the change in

load. The results show that the process is ideally flexible for variable load due to the

characteristics of the thermal recovery section. The performance of the nominal load

design when operating at a given partload matches the maximum performance of the

process designed specifically for that partload. When designing the process specific to

a partload, the turbines are maintained at the nominal load design in order to allow

for a full range of load operations. The ideally flexible behavior is owed to the ther-

mal recovery section, which is not affected by the reduction of the pressure ranges of

the turbine expansion line with decreasing thermal load. The recovery section always

provides adequate preheating to the working fluid. In particular, a unit flowrate of

working fluid always receives the same preheating from the flue gas at the thermal

recovery section independent of the operating load. This signifies a relatively larger

preheating duty with the decrease in the turbine pressure ranges; the recovery section

compensates the decrease in the effectiveness of the inflexible regeneration section,

and therefore, the OCC process is ideally flexible. This flexibility is in contrast to

Rankine cycles without pressurized recovery, wherein the performance significantly

deteriorates compared to the nominal operation. Moreover, as the thermal load and

the working fluid flowrate decrease, the required turbine pressures fall below the crit-

ical pressure, and the working fluid in the HRSG pass through the saturation region.

However, this transition occurs at a larger load for Rankine processes without pressur-

ized recovery compared to the pressurized OCC processes which are able to maintain

supercritical conditions at smaller thermal loads.

Due to the increasing effectiveness of the recovery section, the working fluid enter-

174
ing the HRSG is high enough such that the flue gas of the pressurized OCC process

never violates the acid condensation constraints even at extremely low loads. In

contrast, due to the reduction in the effectiveness of regeneration, conventional coal

Rankine cycles are constrained by the acid condensation temperatures at certain part

loads, and therefore operate sub-optimally with a low performance.

175
176
Chapter 5

A Split Concept for HRSG with

Simultaneous Area Reduction and

Performance Improvement

5.1 Summary

A split concept for boilers and heat recovery steam generators (HRSG), where flue gas

recycling is required for controlling the maximal temperature, is proposed for reducing

the heat exchange area and/or the recycling power requirements. The concept is

demonstrated in the context of an HRSG of a pressurized oxy-coal combustion process,

where the hot flue gas entering the HRSG is diluted by recycled flue gas to comply

with the temperature constraint. The split concept proposes splitting the hot flue gas

prior to dilution, and introducing the splitted fraction, with or without a secondary

recycling stream, at an intermediate point in the HRSG. As a result, the split allows

for lower recycling power requirements (lower diluent flowrate) and a smaller heat

exchange area because the average temperature difference between the hot and cold

streams in the heat exchanger is increased. Multi-objective optimization, for area and

177
power requirements, is performed and the Pareto front is constructed. Results include

a reduction by 37% without a change in the compensation power requirements, or a

decrease in the power requirements by 18% (corresponding to 0.15 percentage points

in cycle efficiency increase) while simultaneously reducing the area by 12%.

5.2 Motivation

Reducing the capital cost and/or increasing the efficiency of power generation is

highly desirable especially given the ever growing market of electric power [36]. Be-

sides economic concerns, efficiency in power generation is also important because the

dominating majority of electric energy production is from non-renewable fossil fu-

els, and there are increasing concerns and regulations regarding their emissions, [37].

Flexibility to uncertain parameters, like fuel specifications, ambient conditions, and

thermal load is another important characteristic required from the implemented tech-

nologies.

In thermal power generation there are several forms of unavoidable losses as well

as several operational constraints and economic considerations. For example, the area

of heat exchangers areas are limited by capital cost consideration, resulting in less

effective heat transfer and/or larger pressure drops of the streams due to the packed

and constrained pathways; thus, the efficiency of the cycle decreases. Metallurgic

properties pose temperature constraints on combustors, boilers, heat recovery steam

generators (HRSG), turbines, etc., [40, 22], thus resulting in exergy destruction and

reduction of power output. For example, flue gas recycling (FGR) in coal boilers is

required in order to quench the combustion gas to limit the radiative-dominant (high

temperature) heat transfer regiment and enables a convective-dominant heat transfer;

FGR is applied because radiation is more expensive than convection for the same

degree of thermal energy transfer, [65]. FGR also increases the boiler efficiency and

178
reduces emissions, and is applied in almost all relatively recent (younger than 30 years

old) coal fired powerplants, [65, 66]. The FGR require compression to compensate for

losses in the boiler and the recycling pipes; moreover, throughout the heat exchange

process, the temperature gradient between the hot and cold streams decreases as flue

gas moves from away from the inlet, which increases the heat exchange area required

close to the cold end of the heat exchanger.

One of the promising concepts of carbon-capture and sequestration is pressurized

oxy-coal combustion (OCC). Pressurizing the flue gas increases the effectiveness of

the convective heat transfer. In [20, 21, 23, 25, 24, 27, 38, 35, 67] and Chapters 2-4, a

pressurized OCC concept is considered with an HRSG with relatively high FGR that

relies on convective heat transfer. Note that in general the combustion process occurs

in a section, that may or may not be physically connected to the HRSG, referred to

as a combustor. The capital cost of the HRSG is a relatively large portion of the

capital cost of the powerplant, e.g., [68], and reducing its size results in significant

savings.

Herein, a novel split concept for the HRSG is introduced in order to enhance the

rate of thermal energy transfer by increasing the average temperature between the

involved streams, and reduce the compression requirements by reducing the recycling

flow rates and/or pressure losses compared to the conventional operation. The concept

is applicable to coal boilers and other heat exchange processes that require quenching

of the hot fluid. Section 5.3 applies the concept to a standalone model of an HRSG of

the aforementioned pressurized OCC. Section 5.4 deals with minimizing the area of

the exchanger and the compensation power requirements while keeping fixed the input

streams conditions and the total transferred thermal energy. A detailed explanation of

the two objectives, variables, and constraints are presented, and the Pareto front of the

multi-objective optimization is constructed. The results are discussed in Section 5.5,

where the Pareto front illustrates the achievable reductions in the area and power

179
requirements, allowing for lower capital costs and/or higher efficiencies. Section 5.6

shows that the split concept preserves the ideal flexibility of the pressurized OCC

process.

5.3 Novel Split Concept

5.3.1 Concept Description

Figure 5-1 illustrates the split concept applied to the HRSG of a pressurized OCC

process, [24, 27, 38, 35, 67] and Chapters 2-4. In pressurized OCC, oxygen is delivered

to the combustor at an elevated pressure. Primary recycling flue gas, FG-Rec-pri,

is mixed with the oxygen stream for dilution in order to control the temperature

of the combustor to acceptable levels, [40]; a temperature of 1550*C is considered

here. Combustion gas is mixed with secondary recycling flue gas, FG-Rec-sec, to

achieve an acceptable temperature at the entry of the HRSG. In the HRSG, thermal

energy is transferred to the working fluid of a Rankine cycle. Flue gas is recycled for

controlling the critical temperature of the combustors and the HRSG in two possible

configurations, wet or dry recycling. In wet recycling flue gas is recycled directly

after the HRSG exit, while in dry recycling flue gas is recycled after condensing and

separating the water. Figure 5-1 illustrates the wet recycling case, but the split design

can be equivalently applied to dry recycling. Recycling fans are used to compensate

for the pressure losses encountered by the flue gas mainly in the HRSG and the

recycling pipes.

The concept proposes splitting the hot combustion gas prior to its dilution by the

secondary recycling stream. The flue gas entering the HRSG decreases in tempera-

ture as it exchanges thermal energy with the working fluid of the Rankine cycle; flue

gas acid condensation in the HRSG is not allowed as discussed later in the operation

constraints, Section 5.4.3. In essence, the mixing of the splitted stream in the HRSG

180
is intended to elevate the flue gas temperature. The primary flue gas is mixed with a

recycling stream similar to the HRSG without splitting. The splitted flue gas is (po-

tentially) mixed with another secondary recycling stream. The secondary recycling to

the split allows larger ranges and larger feasible combinations of the mixing positions

and the mixing temperatures. The specific amount of recycling to the split, if any,

is determined by optimization. The results in Section 5.5 demonstrate that minimal

area does not require any recycling to the split, whereas minimal compensation power

requirements requires. The additional split pipe, the pipe of the recycling to the split

(if used), and the recycling to the split recycling fan (if used) add some capital cost,

however, it is insignificant compared to the savings in the HRSG. Note that even

when two recycling streams are used, a single fan can be installed by introducing

some throttling at the outlet of the split recycling stream; obviously this results in

higher compensation power requirements. For a given HRSG thermal load, the split-

ting process can increase the overall temperature difference between the streams of

the exchanger, particularly avoiding small temperature differences which require the

most heat transfer area. Moreover, the split allows for lower rates of recycled flue

gas compared to regular HRSG: the required amount of recycling to the inlet of the

HRSG is smaller because it needs to dilute a smaller amount of the combustion gas.

Also, at the mixing position, the flue gas within the HRSG acts as a diluent to the

split stream, thus a relatively small amount of the recycling to the split is required

(if any).

The higher average temperature differences imply smaller exchanger area, and

the lower recycling requirements imply lower pressure drops and lower compensation

power requirements (CPR). The CPR consists of two components; the first is the

power needed by the recycling fans to re-pressurize the recycling streams to compen-

sate for the pressure losses encountered in the HRSG and the recycling pipes. The

second component of the CPR is the power needed to maintain the main flue gas

181
FG-Rec-pri
to combustor TMix
FW-HRSG-in FG-R
Reheat
c-s rr 0
FG-Rec-sc T
Rec-FanO

Combustion Split Hot- as


Thermal -Rcvry-out
GsGaso 'Ht a ool-Gas
to CS

Tempe~r.'0 FG-Rcvry nRecovery


plte r
th
91jt ~~Controller WRvynC
split F-cr ondensate

plit -Rec-sec 1
basi
Split-Rec Rec-Fani
TiX Mixerl1c
Mixm ecl
Figure 5-1: HRSG single-split as part of a pressurized OCC process with a thermal
recovery unit. Optimization variables in purple circles marked with (o). Constraints
are on the maximum allowed temperature in the HRSG and safety margins against
acid condensation.

flow as it faces pressure losses while passing through the HRSG, Section 5.4.1. For
simplicity, only one split is illustrated and tested here. Multiple splits, that would

be introduced sequentially at different intermediate locations, are possible and would


further increase performance and decrease the heat transfer area, but would add

structural complexity. Note that in general the split can be extracted from any point
along the HRSG and not limited to the inlet combustion gas, and the recycling can
also be withdrawn from any point along the HRSG and not limited to the outlet Cool-

Gas; however, such processes are less practical to implement and more complicated
to model and optimize.

5.3.2 Stand Alone HRSG-Split Simulation

To illustrate the possible advantages, Figure 5-2 shows the profiles of four cases of

HRSG operation as a standalone unit with an identical set of input parameters.

182
As shown in Table 5.1, the input parameters are the specifications of the hot and

the cold input streams, the outlet specifications of the cold stream, and fractional

thermal losses; the fixed specifications signify that the total amount of thermal energy

transferred in the HRSG is fixed. Table 5.1 shows two sets of input parameters

as obtained from an optimized pressurized OCC with wet recycling, [35, 67] and

Chapters 3&4; the specifications titled CoalA are the operating conditions that the

HRSG encounters when the basecase OCC process utilizes a high quality coal, while

those titled CoalB are relevant to combusting a lower quality coal, where CoalA

and CoalB are identical to those utilized in [35] and Chapter 3. The specifications

titled CoalA are used here upto Section 5.5, and then for the flexibility assessment

(Section 5.6) both coals are used. Both sets of input specifications of the streams

are for the nominal full load operation. The cold stream profile is that of the main

feedwater and the reheat streams. The maximum allowed HRSG temperature is

800*C equal to that considered in [24, 27, 38, 35, 67] and Chapters 2-4. Similar to the

basecase the HRSG is considered to face a 0.75% fractional heat duty losses. First,

note that in Figure 5-2 although the input streams specifications, CoalA specifications

of Table 5.1, and the total duty transfer in the HRSG are constant among all four

operations, the temperature of the flue gas at the exit of the HRSG, Cool-Gas, might

not be. The different pressure losses and recycling requirements for each operation

result in different CPR which causes different amounts of compression enthalpy rise

(CER) carried by the flue gas, [24, 27]; therefore, the temperature of the Cool-Gas

exiting the HRSG is slightly different between the four profiles.

The first profile in Figure 5-2 is without splitting or recycling, i.e., all the com-

bustion gas enters the exchanger directly without dilution; this operation violates the

maximum temperature constraint on the HRSG, and is only shown for illustration.

The infeasible operation of Profilel theoretically requires the smallest heat exchanger

area due to the largest temperature differences between the hot and cold streams, and

183
requires zero recycling and zero second component of the CPR; (the CPR components

are described in Section 5.4.1).

The second profile in Figure 5-2 represents the standard basecase operation where

all the combustion gas is mixed with enough recycling to result in Hot-Gas entering

the HRSG at precisely the maximum allowed temperature. The flue gas temperature

then drops to the exit temperature as thermal energy is transferred to the working

fluid. The flue gas has approximately constant thermal capacity as inferred by the

nearly linear temperature profile versus thermal energy transferred. A lower inlet

temperature for the Hot-Gas into the HRSG requires higher recycling flowrate, which

results in larger flue gas flowrates in the HRSG, a flatter temperature profile of the

hot stream, and smaller temperature differences between the streams of the HRSG. If

thermal energy transfer and pressure losses are independent of the flow conditions in

the HRSG, then lower inlet temperatures, leading to smaller temperature differences

and larger recycling flow rates, are clearly unfavorable regarding both exchanger area

and the CPR. However, as described in Section 5.4, the heat transfer, pressure losses,

and flow conditions of the flue gas are not independent, so larger flows and smaller

entry temperatures might be favorable in some cases, especially when CPR is of a

higher priority than area, since larger flowrates may contribute in smaller HRSG

pressure losses.

The third profile in Figure 5-2 represents a theoretical operation while respecting

the maximum temperature constraint; this graph is only given for illustrative pur-

poses. The profile is achieved by an infinite number of splits, and an infinitesimal

recycling to the inlet required to decrease the temperature of the infinitesimal inlet

combustion gas from 1550'C to 800'C. The splitted combustion gas is introduced

infinitesimally into the HRSG maintaining for as long as possible a constant temper-

ature equal to the maximum allowed. When all the combustion gas is introduced, the

temperature profile is only infinitesimally flatter than that of Profilel, and the two

184
profiles seem indistinguishable.

The fourth profile in Figure 5-2 represents an operation with a single split. First a

certain amount of combustion gas is split, and just enough recycling to the inlet of the
0
HRSG is used to obtain a Hot-Gas temperature of 800 C. The split is then introduced

to the HRSG without any recycling at a point where the resulting flue gas mixture in
0
the HRSG attains a temperature of 800 C. Compared to the conventional operation

of Profile2, the split can provide larger average temperature differences between the

steams of the HRSG, therefore, smaller areas. Moreover, lower recycling flowrates,

and possibly smaller CPR, are required since respecting the maximum temperature

constraints are attained not only by recycling but also by the gradual heat transfer

(note that pressure losses in the HRSG, which might increase, are another factor
in determining the CPR). Lower flue gas flowrates can also be inferred from the

steeper slopes of Profile4 compared to those of Profile2. Introducing the split further

downstream and/or adding recycling to the split, neither of which are shown here but

considered in later sections for optimization, result in a lower mixture temperature

inside the HRSG. Also, adding recycling to the split can allow earlier mixing positions

while satisfying the maximum temperature constraint.

It can be proven geometrically that for a given split flowrate, the largest tempera-

ture differences between the hot and cold streams are attained when the inlet and the

mixing temperatures are at the maximum allowed and when there is no recycling to

the split. Also, by comparing the slopes of the temperature profiles, it can be proven

that for any split flowrate, operating with the mixing temperatures at the maximum

possible value minimizes the flowrate of the recycling streams. It is tempting to say

that for a given split flowrate the least area requirements and the least recycling

requirements are obtained when the constraints on the maximum allowable temper-

atures are active; as the optimization shows, in fact the area is minimized when the

maximum temperature constraints are active, however, this is not always true for the

185
la= Prfll 50 ,iswn=Ogs Mn.NIRW=Ok/,r~,N
Profile1: Tuiw,0 = 1550* C, 7hsyntj = Okg/s, Tmi., =NA, 7hac.0 = Okg/s, 7hjz. =NA
No Split, Constraint Violated

1 400 --
Profile2: TMjO = 800*C, 7hspam = Okg/s, &.,I =NA, hR.,I = Okg/s
Original operation
Profile3: imi.O = 800*C, 7sj, (1-1d) = ekg/S, R,,O = e, he, (,-,of = Ok
1
Infenitesimal Splitting, e recycling to inlet

000 - Profile4: fusj.0 = O*C, hsoit, 150kg/s, mi, = 800*C, rhRcj = Okg/s
Single Split
a
0.
a

400 -

Cold streams (main Feed Water and Reheat streams) temperature profile common
among all four hot stream profiles
2M 1 1 1 1 1
1 2 3 4 5 6 7 8
HRSG Duty Transfer, QHRSC (W) XO'

Figure 5-2: Temperature profiles of four different operations of the flue gas with iden-
tical cold streams profile (and heating duty requirement). Profilel has no recycling or
dilution and violates the maximum temperature constraints. Profile2 is the original
basecase operation. Profile3 has infinite number of split. Profile4 is an un-optimized
example of the flue gas with a single split, where the overall temperature differences
between the streams of the HRSG can be higher than the original operation, and the
recycling flowrates are lower.

CPR.

It also can be proven that any split operation with a certain total amount of recy-

cling, regardless of the number of splits, the splits flowrates, or the mixing positions,

can at most reach the boarders outlined by a profile with no split and the same total

amount of recycling introduced all at once to the inlet; as an example Profile3 and

Profflel with an infinitesimal recycling to the inlet, respectively.

186
Table 5.1: Fixed input parameters for the HRSG. Two different flue gas
conditions are presented, each relevant to a different coal type. The condi-
tions titled CoalA are used up to Section 5.5, and both the conditions of
CoalA and CoalB are used for the flexibility assessment of Section 5.6.

Input Parameter With CoalA With CoalB

Flue Gas Conditions


Combustion gas flowrate 402kg/s 394kg/s
Combustion gas tempera- 155b C
ture
Combustion gas pressure 7.41bar 9.67bar
Combustion gas mole frac- H2 0 = 0.479; 02 H 2 0 = 0.478; 02
tion Composition =0.030; N 2=0.008; =0.030; N 2 =0.009;
C0 2=0.457; C0 2=0.458;
S0 2 =0.001; S0 2 =0.001;
AR=0.025; AR=0.024;
Maximum allowed HRSG 800 0 C
temperature
HRSG fractional heat loss 0.75%
Flue Gas flowrate to recov- 120kg/s 132kg/s
ery unit
Feedwater and Reheat Con-
ditions
Feedwater flowrate 306kg/s 301kg/s
Feedwater inlet tempera- 322 0 C 322.0 0 C
ture
Feedwater inlet pressure 265bar
Feedwater outlet tempera- 600 0C
ture
Feedwater outlet pressure 250bar
Reheat flowrate 233kg/s I 260kg/s
Reheat inlet temperature 358OC
Reheat inlet pressure 53.5bar
Reheat outlet temperature 610 0 C
Reheat outlet pressure 53. ibar
Compressors Specifications
Primary recycling fan misentropic = 0.83
77mecbanical = 0.99
Thermal spec = Adiabatic
Secondary recycling fanO&1 nisentropic = 0.8338
7
7mechanical = 0.99
Thermal spec = Adiabatic
Main stream compensation 77isentropic 0.90
compressor ?7mechanical = 0.98
Thermal spec = Adiabatic
5.4 Optimization Formulation for Minimal Area

and/or Minimal Compensation Power Require-

ments

5.4.1 Objective Functions

The purpose of the split is to reduce the capital cost by utilizing smaller surface

area of the heat exchanger and reduce the CPR by reducing pressure drops and/or

recycling flowrates. The two objectives are neither equivalent, nor mutually exclusive,

and have to be accounted for simultaneously; smaller exchangers in general lead to

larger pressure losses, and the flow properties affect the heat transfer coefficient, the

HRSG pressure drop, and the recycling losses. Herein, a multi-objective approach is

taken and the Pareto front, or set of non dominated solutions is obtained. This is

constructed using a weighted sum approach, [69], and hierarchic optimization, [44, 43],

as explained later.

Objective Functions Calculation

Computing the objectives is not trivial and herein, they are based on similarity anal-

ysis for calculating the area and pressure losses.

Area of the Heat Exchanger

The HRSG area is calculated according to the logarithmic mean temperature differ-

ence, and discretized heat exchanger (1000 elements). For each discretized region, i,

we have:
Aa,i oa,i Ub,i Fb,i ATLM, b,i
Abi Qbi Uai Fai ATLM, a,i

where subscripts a and b stand for actual and basecase respectively. A is the area,

U is the equivalent heat transfer coefficient, Q is the heat transfer duty, and F is its

188
correction factor.

Compensation Power Requirement, CPR

The CPR consists of two components: the first component of the CPR is the power

needed by the recycling fans to re-pressurize the recycling streams after experiencing

pressure losses from the HRSG and the recycling pipes. The second component is the

power needed to maintain the main gas flow and overcome the HRSG pressure losses.

The extra compression needed to overcome the main flow losses can be introduced

prior to combustion or after the HRSG. For example, in standard coal power plants,

the same fans or compressors that drive the inlet streams to the combustor force the

flow through the main stream pressure losses. But in pressurized OCC combustion,

since a compressor is already present after thermal recovery and needed for the carbon

sequestration unit (CSU), the compensation power requirement is less costly to be

accounted for by compressing a cooler stream with a lower flowrate post the thermal

recovery.

Similar to the basecase, the pressure loss of each recycling pipe, APpjp, is esti-

mated by the following equations [62, 63]:

2
APipe = pf L V

where V is the bulk gas velocity in the pipe, d is the diameter of the pipe, Lp is

the equivalent length of the pipe, p is the gas density, and f is the friction factor

calculated by

- -2
(2e/d) _5.02 (2e/d) + 13
13
\
fpipe 2.-20Olog (2~) .2log
7.4 Red 7.4 Red

where c is the pipe roughness, Rea PVd is the Reynolds number based on the pipe

diameter and M is the dynamic viscosity of the gas. The pipe diameter, d, and the

189
gas velocity, V, are related by n = pVd, where fr is the recycled gas mass flowrate

through each pipe. The diameters, d, and equivalent length of the each recycling pipe,

LP, are identical to those of the basecase, optimized results of [24, 27], where practical

and economic considerations and experimental data are incorporated in obtaining the

diameters and the equivalent length; therefore,

APipe,a Pa fa Lp, adb _ faaKPb


APpipe, b 2
Pb fb Lp, b da V fbaPa

The subscripts a and b stand for actual and base-case, respectively. For deriving

the final equality in the above equation the gas mass flowrate is ii = L. Each split

adds a secondary recycling pipe and therefore for the single split considered here two

secondary recycling pipes are present, one to the inlet of the HRSG and the other

to the split at some intermediate point in the HRSG. Both secondary recycling pipes

are considered to have the same specifications, which is a conservative approximation

since the recycling pipe to the split may be shorter than that to the inlet.

The primary recycling is also considered but treated a little differently in order to

preserve the inlet combustion gas conditions for the stand alone model. The primary

pipe losses and compression requirements are evaluated for a free-open-end stream

without reintroducing that stream into the combustor that outputs the combustion

gas, which is considered as a fixed input parameter. This is intended to maintain

the inlet stream conditions of the stand alone model and reduces the simulation

complexity by eliminating the need to converge an additional outer-most recycling

stream. The pressure losses and power requirements of the free-open-end stream are

very close to those of a closed loop stream, less than 1% difference. Similarly, the

properties of the flue gas are not significantly different with the open-ended stream

in the standalone model compared to those when including the combustor. The

basecase recycling conditions are: APpd


1 ,b = 0.058bar, APse, b = 0.035bar, Tpri, b =

190
281.lkg/s, hsec, b = 736.5kg/s, and the density of the flue gas at the exit of the

HRSG/inlet to recycling pipes is pb = 4.47kg/M 3

The pressure drop in the HRSG is more complicated to evaluate and is expressed

as, [62, 63]:

V2
= Npf max
APHRSG
2

where Vmax = V S$D is the maximum velocity between the tubes and V is the aver-

age face velocity [62]. N is the number of tube bundles/rows along the longitudinal

direction, and the constant parameters D and ST are the tube diameter and the

transverse pitch of the fixed design heat exchanger, respectively. The friction factor,

f, is a function of Reynold's number. All operating conditions considered herein re-


sult in high Reynold's numbers and thus an approximately constant friction factor.

Therefore,
APHRSG, a NaPa 0 a (5.1)
APHRSG, b NbpbV (5.

N is directly and linearly proportional to L, the length of the heat exchanger, by

a factor of 1/SL, where SL is the longitudinal pitch and is also constant for a fixed

design.

The heat exchanger surface area is A. = irD x W x ST


x -S,
SL
where H is the

height of the exchanger and H/ST represents the number of tubes in the transverse

direction. W is the width of the exchanger which is also the length of the tubes

through which water/steam pass. Therefore A. = constant x AcL where A, = x 2 is

the HRSG cross sectional area when, for simplicity, a square cross section of side x

is assumed (H = W = x). Also Vo0 = A. r= xcpA. stant, therefore


threor ,,Vb =
7rnbpaAs, aLb

which implies
La _rbPaAs, aV, a
Lb - raPbAs, bVo, b

191
and as a result
APHRSG, a As, aP Vc0 arnb
APHRSG, b As, bPb 0 , bra

The Reynolds number based on the hydraulic diameter Dh is given by ReDj =

PVOD, where Dh = -Ae. For the square cross sectional area considered,
p Weted Peremeter

Db = x = A=h C(M
VOP )1/
2
which means that ~~ReDh
Reh " - Vi
,b VbP 0,

Reynolds number is approximately related to Nusselt number by ReDh, _ ,a


ReL, b N Ih~
where Nu = hDh. h and K are the gas heat transfer coefficient and thermal conduc-

tivity respectively. h is usually the limiting factor in U and therefore for simplicity

can be considered comparable. Moreover, the other resistances of U are constant due
to the fixed parameter specifications of the stand alone model. Usually a = 0.8, [62].
This leads to
VO, 2 _hc'kbanaSL1 a+1
2PaPb
VO, b hkaarhb 2LbPa 2

substituting in the pressure drop equation

APHRSG, a As, a ha) 60 Kb) (Ila)2 6a (-a")


APHRSG, b As, b (ha Ka \j#b/ \nb/ \PPa/

Where APHRSG, b=0.265bar. Two simplifying and justifiable approximations can

be used. Introducing the split does not require changing the internal design of the

heat exchanger. Therefore, the heat transfer coefficient on the gas side, h, and the

equivalent heat transfer coefficient, U, can be considered equal between the actual

and the basecase operations. Second, the logarithmic mean temperature difference

correction factor F, can also be considered constant. F is a function of the design

and the temperatures at the extremities of each discretization, but in the range of the

optimum temperatures obtained in the results, the correction factor is comparable to

that of the basecase. In fact, Fa is smaller than Fb prior to introducing the split so the

approximation of Fa = Fb contributes in an underestimation of the actual area prior

192
to mixing the split, but larger after introducing the split so the same approximation

contributes in an overestimation of the actual area post mixing of the split. Since

the optimization results in Section 5.5 show that the split is introduced closer to

the inlet than to the outlet of the exchanger, then the assumption of constant F is

conservative. Finally we can write the objective functions as follows:

Minimize the ratio of the actual area to that of the basecase:

1000 Aa,i Aa,i - a,i ATLM, b,i


mi Ab Ab,i Qb,i ATLM, a,i

and minimize the ratio of the actual compensation power requirements relative to

that of the basecase:


. CPRa s.t.:
mm
CPRb

CPRi = nRec-pri (h(TFRec-pri, PComb-Gas) - h(TCool-Gas, PComb-Gas - APHRSG - A pipe pri))i


CPRpri, i

" Rec-sec0 (h(TFG-Rec-seco, PComb-Gas) - h(TCool-Gas, Comb-Gas - APHRSG - APpipe secO))i


CPRseco, i

± ThRec-seci (h(TFG-Rec-sec1, PFG-HRSG) - h(Tool-Gas, PComb-Gas - APHRSG - APpipe seci


CPRs.c, i

± mRecov-out (h(TCo2-OUt, PCO2-0 t) - h(TRecov-out, PComb-Gas - APHRSG)).


CPR main flow, i
i E {a, b}

where h is the enthalpy of the flue gas. Note that the temperatures of the streams

exiting the compressors,TFGRec-pri & TFG-Rec-sec & TFG-ReC-seC1 & TCO2-0 1t, are depen-

dent variables and defined by the respective streams entering each compressor and

the characteristics of the compressor. The pressure losses are:

2
APpipe, a pa fa Lp, a db V fa Tl Pb
APpipe, b Pb fb Lp, b da Vb fb 72Pa

193
and

6a 6 2(a-2)

APHRSG, a _ As, aK "*a +1Tha"a+1 Pb


6Ha 6 2(a-2)
APHRSG, b As, bKa".' 1" Ak+
h a
ba [b mb Pa

5.4.2 Optimization Variables

As aforementioned, a single split is considered herein. The independent decision vari-

ables are chosen to facilitate optimization, since they allow satisfying some constraints
by properly setting their ranges as shown in Section 5.4.3. Also, these variables are
considered the simplest to monitor and set to their desired values during operation.
The optimization variables are: (i) the split flowrate, riiht, 1 , (ii) the temperature at

the inlet of the HRSG, TMiX, 0, (iii) the flowrate of the recycling stream to the split,

7hRp, 1, (iv) and finally the temperature of the flue gas in the HRSG after introducing
the split, TMiX, 1. The variables are illustrated in Figure 5-1 and the ranges are defined

in Table 5.2.

Based on this choice of independent variables, important variables are now de-

pendent. For example, the flowrate of the recycling stream to the inlet, hRc, 0, is

dependent once the split flowrate and the inlet temperatures are specified; i.e. the

stream entering the HRSG is fully specified. Further, the position of introducing
mixture of the split and its recycling in the heat exchanger, Mix-Pos= befre mxng is

dependent once the split flowrate, the flowrate of the recycling to the split, and the
mixing temperature are specified.

5.4.3 Optimization Constraints

The operation of the heat exchanger is subject to physical limitations. Introducing


the split provides better performance while still satisfying these constraints. The

temperatures of the streams inside the heat exchanger are a major concern. De-

194
Table 5.2: Optimization variables, their ranges and the basecase default values,
for a single split. Because most of the basecase variables values are far from the
optimum (zero split, no recycling to split, and no mixing within the exchanger),
several initial guesses are implemented in order to exclude suboptimal convergence.
The boundaries of the ranges of the temperature variables are set to avoid con-
straint violations. TMiX, o lower bound is set to the maximum temperature of the
cold streams, i.e., the reheat stream exiting the HRSG at 610'C; also, the upper
bound on mixing temperatures, TMix, O & TMix, 1, are set to the maximum allow-
able temperature in the HRSG 800'C; and the lower bound of TMiX, 1 is set to the
temperature of feedwater entering the HRSG 321.7 C.

Number Variable Range Base-case and/or default value


1 rnSpliti [0 - 300] kg/s 0/set to 100 kg/s
2 TMix, o [TFW, out - THRSG, max] 'C THRsG, max = 800'C
3 m ec, 1 [0 - 600] kg/s N/A /0
0
4 Tuix, 1 [TFW, in - THRSG, max] C N/A set to THRSG, max = 800'C

fined by the metallurgic properties, the maximum temperature allowed in the HRSG

is limited to THRSG, max = 800'C. In essence, the constraint has to be satisfied at

every point within the heat exchanger; but because the temperature monotonically

decreases along the HRSG (apart from the mixing point), the constraint needs only

be imposed at the inlet and at the mixing position. For the addressed values of the

input streams, the temperature of the cold stream entering the HRSG is safely above

the acid condensation temperature of the flue gas, [24, 27], therefore, there is no need

to include constraints on the minimum allowed temperature for avoiding condensa-

tion of acids in the flue gas or on the feedwater tubes in the HRSG. In contrast,

it is beneficial to include constraints on the MITA to avoid temperature crossovers

and ensure a realistic operation. The physical limit on MITA is zero, but a value of

0.5'C is used to speed the optimization process. In other words, the intuition that

small MITAs are clearly undesirable since they result in huge area requirements, is

communicated to the optimizer. The results show that the constraint on MITA is not

active along the Pareto front and thus does not limit the minimum CPR operations,

195
or the Pareto front profile.

5.4.4 Pareto Front Construction

The two objective functions are dependent and accounted for simultaneously by find-

ing the Pareto front curve. Optimization is performed by combining the two objectives

in a weighted sum approach with modifications. First, two independent optimization

runs are performed to determine the value of the minimum area ratio, min A and

minimum CPR ratio, min 8 , respectively. Then, two hierarchic optimization runs,

[44, 43], are performed to determine the two ends of the Pareto front, Point A and

Point B. Point A is a result of the hierarchic optimization that minimizes CPRp


CPRb
sub-

ject to minimal area Aa. Point B is due to that of minimizing L subject to minimal

CPRa. The additional constraints are imposed with a tolerance of 0.01. Afterwards,

multi-variable optimization is performed using the weighted sum approach combining

both objectives in order to construct the intermediate points of the Pareto curve.

Fifty steps are used in the weighted sum, each with ten multi-start points. Figure 5-3

shows the Pareto front of the HRSG-split where the x-axis is the ratio of the actual

area to the basecase area, , and the y-axis is the ratio of the actual CPR to the

basecase CPR, CPR.


CPRb For validation, 5000 additional simulation runs picked up from

a grid spanning the variable space are evaluated and 225 points close to the Pareto

front are shown in blue dots; all lie to the right and above the Pareto curve. The orig-

inal no-split/zero-split flowrate basecase operation has the coordinates (1, 1) where

CPRbase = 8115kW.

5.5 Results

The Pareto front clearly shows the tradeoff between the two objectives: a small

surface area results in a large pressure drop due to the flow constrictions. Starting

196
1.25 I I I i I I I

1.2

1.15

W0 1 .1 Point A

U 1.05F -7. basecase

2 0 95F
1

0.9
- *Point B
0.85

0.8
0.7 0.8 0.9 1 1.1 1.2 1.3 1.4
Aactual/Abase

Figure 5-3: The ratio of the compensation power requirements versus the ratio of the
areas, both relative to the basecase operation. The Pareto front, obtained using dis-
cretized weighted sum approach, multi-start optimization in each step, and hierarchic
optimization, of the single-split HRSG is shown in the red dashed line. Additional
225 simulation runs are plotted in blue points. The basecase power requirement is
8.1MW for a net electric power output of 300MW.
from point A of coordinates (0.63, 1.05), as area increases the compensation power

requirements decrease rapidly at first since in the region of small areas the HRSG

pressure drop is the predominant form of loss. For further increase in the surface

area, the CPR decreases at a much smaller rate and becomes almost insensitive to

the area changes. In the range of large surface areas the recycling pipes pressure

losses are predominant, and therefore the CPR are not noticeably effected by the

HRSG size. Beyond an area ratio of 0.88 the CP cannot be reduced any lower than

0.82 leading to point B of coordinates (0.88, 0.82). The range of variation in the

CPR for this HRSG contribute around 0.15 percentage points in the efficiency of the

powerplant. Larger savings in area and in CPR can be encountered depending on the

input parameters and operating conditions, as explained in Section 5.7.

The detailed results of the two ends of the Pareto front, points A and B are shown

in Table 5.3. Designing for the minimum area, Point A, requires the largest split

flowrate of all designs on the Pareto curve, rhSpiit, = 180kg/s, requires zero recy-

cling to the split, mhR,,ci, and mixes the furthest away from the inlet, where the Mix-

Pos=Abefore mixing =0.173; the mixing position along the Pareto curve is always closer to

the inlet than to the outlet, therefore, the approximations considered in Section 5.4.1

are conservative. The large split ensures that the temperature of the flue gas within

the HRSG and at a point relatively far from the inlet is increased to the maximum

allowed, so that most of the HRSG is operating with large temperature differences

between the hot and cold streams. Dilution by recycling is minimal in order to main-

tain the largest temperature gradients, but results in low flue gas flowrates leading to

the largest HRSG pressure drop, and therefore the largest CPR. For higher area on

the Pareto front, the split flowrate decreases and mixes closer to the inlet. Lower split

flowrate requires higher recycling to the inlet which increases the flowrate of the flue

gas in the HRSG and reduces the HRSG pressure drop. The area increases because a

smaller portion of the HRSG is operating with large temperature differences between

198
the hot and the cold streams, while the CPR decreases because a larger portion of

the HRSG is operating with larger flue gas flowrates. With further emphasis on the

CPR, the recycling to the split increases in order to increase the flue gas flowrate in

the HRSG. Note that the recycling to the split requires just enough compensation

power to achieve the pressure of the flue gas at the mixing position which is slightly

less than that of the HOT-Gas entering the HRSG at the inlet, and this is why the

mixing position is close to but not exactly at the inlet of the HRSG. On the Pareto

front, Point B has the smallest split flowrate, the smallest mixing position, and the

largest recycling to the split.

For this specific study, all the Pareto optimal designs and operations have both

mixing temperatures at the upper bound; i.e., the constraints on the maximum HRSG

temperatures are active. Larger mixing temperatures result in larger temperature

differences between the stream of the HRSG and in lower recycling flowrates, helping

minimize both objectives. This is in general not true; based on the similarity analysis,

the HRSG pressure drop is inversely proportional to the flue gas flowrate. Other

values of the input parameters of the HRSG-split result in Pareto optimal points that

do not have the maximum temperature constraints active particularly at large weights

for the CPR,; increasing the flowrate of the flue gas in the HRSG, mha, by increasing

dilution from recycling, reduces the pressure losses in HRSG at the expense of the

streams' temperature difference, thus reduces the CPR at the expense of the area.

These results are not shown here for conciseness.

5.6 Flexibility to Uncertainties

The operation of a power plant is subject to several uncertainties, and without op-

timizing for a flexible design the performance of the process can suffer significantly,

[35, 67] and Chapters 3&4. The basecase of this study, pressurized OCC process

199
Table 5.3: Optimization results for the operating conditions of streams resulting from the
combustion of CoalA

Variable Basecase Design A Design B

1 0.63 0.88
C 1 1.05 0.82
CMh
Independent Variables
nSplit1 0kg/s 180kg/s 91.8kg/s
TMix, O 800 0C 8000C THRsG, max 800 0C THRSG, max
rnRec, 1 NA Okg/s 169kg/s
TMix, 1 NA 800'C = THRsG, max 800 C THRSG, max
Key Dependent Variables
rhRec, o 737kg/s 411kg/s 570kg/s
Mix-Pos ( m")
mibU"o )
ATOtal
NA 0.173 0.003
APHRSG, before mixing NA/Obar 0.119bar 0.001bar
APHRSG, after mixing 0.265bar= APHRSG, total 0.321bar 0.231bar
APpri pipe 0.058bar 0.055bar 0.054bar
APsec pipeO 0.035bar 0.011bar 0.013bar
APsec pipel NA NA/Obar 0.001bar

utilizing the standard HRSG design, is ideally flexible to fuel specifications and ther-

mal load, [35, 67] and Chapters 3&4. It is demonstrated here that the HRSG split

concept also has this favorable property. A change in coal is addressed here because

among the uncertainties mentioned, the variations in coal have the largest effect on

the streams of the HRSG. Therefore, designing for coal flexibility seems most chal-

lenging. Also, design for variable load flexibility pertains more to the expansion line,

regeneration, and heat recovery sections rather than the HRSG.

A coal powerplant utilizes different types of coals during its lifetime. The advan-

tages of the split concept are not limited to the type of coal used; for any coal utilized

for power generation, the HRSG-split can be designed accordingly in order to reduce

the area and/or the CPR. Since during the operation of the powerplant the utilized

coal is expected to change, it is extremely important that the optimal design of the

HRSG-split for one coal is flexible to changing the coal; i.e., the design is also optimal

200
for the other coal.

Assessing the flexibility of the HRSG-split in general requires characterizing the

variables as design or operation variables, [35, 67] and Chapters 3&4. Design vari-

ables are fixed upon design while operation variables can change with the different

operations of the HRSG. The four independent variables chosen in Section 5.4.2 for

the optimization of the standalone HRSG-split, TMIxo, Thspliti, TMIxi, & rhRecl, are

all operation variables. However, dependent variables of the standalone model, in

particular the split mixing position, Mix-Pos, and the resulting HRSG area, A, are
design variables and have to be common between the different operations. Luckily,

based on the following approach, there is no need to reformulate the problem to in-

clude design variables as decision variables, which pose a lot of difficulties in solving

for the objective function which would require complicated numerical methods, and

result in large domains of infeasible operations.

The ideal flexibility is demonstrated herein by a much simpler approach; the same

optimization above is performed on a standalone HRSG-split but with the a new

specifications of the input streams which are relevant to a coal different from the

original. The input streams specifications used above are a result of the pressurized

OCC process designed for ideal flexibility to uncertainties when operating with a

typical bituminous coal with composition similar to Venezuelan and Indonesian coals

(referred to as CoalA), while the new streams' specifications result from operating
with a lower quality south African coal almost identical to Douglas Premium or

Kleincopje coal (referred to as CoalB), as presented in [35] and Chapter 3. The multi-

objective optimization of the HRSG-split operating with the streams conditions of

CoalB results in a Pareto front curve very similar to that seen for HRSG-split multi-

objective optimization (presented above) operating with the conditions of CoalA.

More specifically, equal areas on the two Pareto fronts have identical mixing positions,

therefore, a given Pareto optimal design of the HRSG-split for one coal is also a

201
Pareto optimal design for the second, therefore, the HRSG is thermodynamically

ideally flexible. Moreover, equal areas between the original and new Pareto curves

are a result of equal weight vectors for the multi-objective optimization, therefore,

the HRSG-split is also economically ideally flexible; determining the most profitable

design does not require to consider the coal distribution because any optimal design

for one operation is optimal for the other, and has the same tendencies/preferences

towards each of the two objectives. A similar behavior to that of the coal variation

is encountered for the other uncertainties. As a conclusion, the HRSG-split is ideally

flexible to uncertainties, and at least capable of maintaining the flexibility of the

process it is incorporated in.

5.7 Other Applications

The application of the HRSG-split is not limited to the OCC process. The split con-

cept can be applied to any heat exchange process that requires a recycling stream to

control the temperature of the main stream; for example, in conventional boilers, both

the FGR rates and the heat exchange area, particularly radiative, can be reduced.

The concept can be readily applied to new power plants; moreover, the retrofit of

existing plants is conceivable.

Although not shown here, the benefits of the HRSG-split in subcritical power

cycles can have larger magnitudes than those obtained here for a supercritical. Recall

that the smaller the temperature differences between the streams, the larger the area

required for the same amount of heat transfer, and therefore, the exchanger area is

directly related to the value of the MITA. In the case study above, the feedwater

conditions are those from an optimized supercritical Rankine cycle, where the pinch

in the HRSG is located at the cold end; the feedwater temperature entering the

HRSG is relatively large, by utilizing the FWHs regeneration, allowing higher rates

202
of feedwater through the HRSG while respecting the MITA specifications, and thus

larger flowrates through the expansion line to increase the power output and efficiency.

Since the pinch is at the cold end, introducing the HRSG-split cannot avoid the pinch

because the flue gas temperature at the exit of the HRSG varies only slightly due to

the variations in the CPR. Also, the feedwater temperature profile in the HRSG is

smooth due to the absence of phase transitions. On the other hand, with subcritical

feedwater, the transition from the subcooled liquid state to the two-phase state is

marked by a sharp kink, and usually the pinch point occurs at that location rather

than at the cold end. Since in an HRSG-split the temperature of the flue gas after

mixing is larger than that of the basecase operation, except very close to the exit of

the HRSG where the temperature might be slightly lower depending on the CPR,

then the temperature difference at the location of the pinch is larger than that of the

basecase. Now since in subcritical operations the pinch is alleviated, the reduction in

area and pressure losses are significantly larger compared to the supercritcal scenario.

Note that the increase in the temperature approach due to the split in the subcritical

process allows, upon process optimization, for even larger feedwater flowrate through

the HRSG which increases the power output and the efficiency; i.e., compared to

the supercritical scenario, in a subcritical process the savings on area and CPR are

larger, and there is a possibility of increased power output and further increase in the

efficiency.

5.8 Conclusion

A new split concept applicable to heat exchangers that require recycling of the hot

stream for temperature control, e.g., coal boilers and HRSGs, is presented herein.

The concept proposes splitting the hot stream, which has a temperature higher than

that allowed in the heat exchanger, before its dilution and its introduction into the

203
heat exchanger. At the inlet of the flue gas, a lower amount of dilution is required

to control the temperature of the now smaller fluid flowrate. The splitted fraction

is then introduced into the heat exchanger at an intermediate point downstream,

increasing the temperature of the hot stream and enhancing the temperature gradient

of the heat exchange process. The concept is able to reduce the cost by reducing the

area requirements and/or increase the efficiency by decreasing the power required to

compensate for the pressure losses of the flue gas. The concept is illustrated in a

standalone model of an HRSG in the context of a pressurized oxy-coal combustion

process. Multi-objective optimization is performed by constructing the Pareto front

of minimal area and minimal power requirements. Both the heat exchange area

and the compensation power requirements are shown to be reduced compared to the

conventional operation; in the illustrated case, the area can be reduced down to 0.63

the original size, and the compensation power requirements can be reduced down to

0.82 the original requirements. The design and operation is not limited to new heat

exchangers and retrofitting is considered easily possible because no changes in the

internal structure of the heat exchanger is required.

Moreover, facing uncertainty in input parameters and operating conditions, the

split concept is shown to be ideally flexible and preserves the flexibility of the process

it belongs to.

Herein, the heat exchange process is enhanced by a slight modification to the

design of heat exchanger while holding the input streams and the total transferred

thermal energy constant. However, the input streams to the exchanger are variables

of the process it belongs to. Therefore, the overall performance of the process and

the performance of the HRSG and the capital cost savings can be enhanced by simul-

taneous optimization of the HRSG-Split and the powerplant deign.

204
Appendices

205
206
Appendix A

Reaction Chemistry Added to the

Separation Column in the DCSC

flowsheet

A set of reactions are implemented in the Separation Column chemistry in order

to properly evaluate the behavior of the condensates in involved streams. Of the

elaborate and relatively large set of possible reactions, only a few are show to be

influential to the scope and assessments of the current work. Table A. 1 presents those

important reactions where their equilibrium constants are evaluated using Gibbs free

energy.

207
Table A.1: The relevant reaction added to Separation Column of the DCSC flowsheet

Reaction type Stoichiometry


Equilibrium 2 NO 2 N20 4
Equilibrium SO 2 + NO 2 - SO 3 + NO
Equilibrium SO 3 + H 2 0 H 2 SO4
Equilibrium HSO- + H2 0 SO2- + H3 0+
Salt Na 2 SO 4 2 Na++ SO-
Salt Na 2 SO 4 .(H2 0)10 - 2 Na++ SO- + 10 H 2 0
Salt Na 2 SO 4 .NaOH 3 Na + SO2- + OH
Appendix B

DCSC Recirculation Water,

rnRW-Sep-in, Optimality Criterion

The optimization of the DCSC flowsheet can be simplified by manipulating the re-

circulating water flowrate to obtain a balanced DCSC-HX. The criterion is justified

after two observations. First, the temperature of the bottom stage of the separa-

tion column, which is equal to the temperature of the recirculating water exiting

the column (RW-Sep-out), is highly insensitive to the flowrate and the temperature

of the recirculating water entering the top stage (RW-Sep-in). For example a 50%

change in flowrate, and/or temperature, in 'C, of the RW-Sep-in results in less than

a 0.01% change in the temperature, in 'C, of the bottom stage/the temperature of

RW-Sep-out. The bottom stage temperature is almost only dependent on the flue

gas entering the separation column (FG-Sep-in) conditions particularly pressure and

constituents because they define the temperature at which water in the flue gas is

transitioning from vapor to liquid. Second, the temperature of the top stage which is

equal to the temperature of the flue gas exiting the separation column (FG-Sep-out),

decreases with the decrease in the temperature and/or increase in the flowrate of

RW-Sep-in; the effect of the RW-Sep-in temperature on the temperature of the top

209
stage is significantly larger than that of the stream's flowrate. On average, within

the practical ranges of operation, a 10% change in TRw-Sep-in, in 'C, results in around

5% change in the top stage temperature, in 'C, while a 10% change in RW-Ser-in,

in kg/s, results in a 3% change in the stage's temperature, in 'C. Analyzing the top

stage temperature is needed since the amount of thermal energy transferred from the

flue gas to the recirculating water and the amount of the water condensed increase

with the decrease in the temperature of FG-Sep-out (and vice versa).

The criterion proposed herein is to have the thermal capacity rate of the recircu-

lating water entering DCSC-HX, i.e., stream RW-HX-in, equal to that of the working

fluid at that section (WF-HX-in), thereby obtaining a balanced heat exchanger. The

specific thermal capacities of the two streams are almost identical and neither are sig-

nificantly affected by the temperature change encountered in the DCSC-HX, there-

fore, the principle is equivalent to matching the flowrates of the streams involved.

The criterion is proven by examining the flowrate and temperature of the splitter's

excess condensed water which is rejected outside the recirculating loop (Split-E-CW),

where a lower temperature and/or a larger flowrate of Split-E-CW signify larger re-

covery and thus better performance. The proof is performed in two parts; first, the

case where i2RW-HX-in < nFW-HX-in is considered and it is shown that an increase in

nRw-HX-in gives an increase in the recovered thermal energy and thus an improvement
in performance; second, the case where rhRW-HX-in > mFW-HX-in is considered and it

is shown that a decrease of nRW-HX-in is favorable. For each case, four general sce-

narios might occur, i.e., the flowrates of RW-Split-in and Split-E-CW can increase or

decrease. It will be shown that the flowrate of RW-HX-in should increase when it is

lower than that of WF-HX-in, and should decrease when it is higher.

First, in the region where the mRW-HX-in is less than hFW-HX-in, the pinch point

occurs at the cold end of the heat exchanger; recall that an initial criterion for op-

timality is for the heat exchanger to operate at the allowed MITA. In this domain,

210
increasing the rnRW-HX-in does not change the temperature of the recirculating wa-

ter exiting the DCSC-HX i.e., the temperatures of Split-E-CW & RW-Sep-in. As

rnRW-HX-in increases, the following four scenarios are conceivable, but only the last is

possible:

1. Both the flowrate of Split-E-CW and of RW-Sep-in decrease; this scenario vio-

lates the mass balance of the splitter and thus is not possible.

2. (&3.) The flowrate of Split-E-CW increases (decreases), and the flowrate of RW-

Sep-in decreases (increases); this scenario leads to a contradiction. A decrease

(increase) in rnRWSep-in increases (decreases) the separation column top stage

temperature resulting in a higher (lower) temperature and water fraction of

FG-Sep-out. This means a smaller (larger) amount of water condensation which

contradicts with the larger (smaller) rhsplit-ECW-

4. Both the flowrate of Split-E-CW and of RW-Sep-in increase; this is the only

remaining scenario out of the four general combinations and doesn't lead to

any contradictions. The increased flowrate of RW-Sep-in, while having the

same temperature, increases the amount of condensed water in the separation

column which is in agreement with the increase in mSpjit-E-CW.

Second, in the region where rnRW-HX-in > rnFW-HX-in, the pinch point occurs at the

hot end of the heat exchanger. Decreasing the flowrate of RW-HX-in decreases the

temperature of RW-Split-in, Split-E-CW, and RW-Sep-in. Again, four scenarios are

conceivable, but only one is possible:

1. Both the flowrate of Split-E-CW and of RW-Sep-in increase; this scenario vio-

lates the mass balance of the splitter and thus not possible.

2. The flowrate of Split-E-CW decrease, and the flowrate of RW-Sep-in increases;

this scenario leads to a contradiction. An increase in hRW-Sep-in and a decrease in

211
its temperature reduces the temperature of the separation column's top stage.

Therefore, the temperature and water fraction of FG-Sep-out decrease. This

means larger amount of water condensation in the separation column which

contradicts with the smaller amount of Split-E-CW.

The two remaining scenarios require first to show that TRW-HX-Out is more sensitive

than rnRW-HX-out with respect to a change in the flowrate of RW-HRSG-in. First

consider the energy balance on the hot stream of the DCSC-HX neglecting pressure

effects:

nRW-HX-inCp (TRW-HX-in - TRW-HX-out) : QDCSC-HX

Consider a fixed cycle operation and thus a constant QDCSC-HX- With an abuse of

notation, differentiate relative to TiRW-HX-in, then multiply by drhRW-HX-in to get:

dIRW-HX-inCp (TR-HX-in - TRW-HX-out) + rRW-HX-incp d(TRW-HX-in - TR-HX-out) = 0

Assuming a constant c, and rearranging:

dhRW-HX-in _ d(TRW-HX-in - TRW-HX-out)


mRW-HX-in (TRW-HX-in - TRW-HX-out)

keeping in mind that TRW-HX-in is equal to the temperature of the bottom stage of the

separation column, which as discussed is independent of the flowrate of RW-Sep-in.

Finally rearrange and divide by TRW-HX-out:

dTRW-HX-out _ TRW-HX-in - TRW-HX-out drhRW-HX-in


TRW-HX-out TRW-HX-out mRW-HX-in

Considering that all temperature are given in 'C, TRW-HX-.Out is less than (TRW-HX-in - TRW-HX-out),
so the percentage change in the temperature, in 'C, of RW-HX-out is larger than the

percentage change of its flowrate, in kg/s (rhRW-HX-in = TnRW-HX-out).

212
3. Both the flowrates of Split-E-CW and of RW-Sep-in decrease; this scenario leads

to a contradiction. The percent decrease in TRW-Ser-in (equal to TRw-split-in/TRw-HX-out

is larger than the percent decrease in rnRWSep-in, with the former being more ef-

fective in changing the temperature and vapor fraction of FG-Sep-out than the

latter as discussed above. Therefore, in this scenario, the temperature and water

fraction of FG-Sep-out decrease. This means larger amount of water condensa-

tion in the separation column which contradicts with the smaller Split-E-CW;

thus this scenario is not possible.

4. The flowrate of Split-E-CW increases, and the flowrate of RW-Sep-in decreases;

this is the only remaining scenario and doesn't lead to any contradictions. The

decrease in TRW-Sep-in reduces the temperature of FG-Sep-out despite the for-

mer' s decrease in flowrate (percent change of flowrate is singificantly smaller

than the change in temperature and has significantly smaller effect even for the

same percent change). Therefore, the amount of condensed water from the flue

gas increases, which is in agreement with the increase of Split-E-CW.

In conclusion, the largest amount of thermal recovery occurs when the stream flowrates

at the DCSC-HX are matched.

213
214
Bibliography

[1] A. Bejan, Thermal Design & Optimization, John Wiley & Sons, Inc., 1996.

[2] R. W. Haywood, Analysis of Engineering Cycles, 3rd Edition, Elsevier, 1980.

[3] M. J. Moran, H. N. Shapiro, Fundamentals of Engineering Thermodynamics, 6th

Edition, John Wiley and Sons, 2007.

[41 Thermoflex@, www.thermoflow.com.

[51 Gatecycle@, http: //www.gepower.com/prod-serv/products/oc/en/

opt-diagsw/gatecycle.htm.

[6] L. T. Biegler, I. E. Grossmann, A. W. Westerberg, Systematic Methods of Chem-

ical Process Design, Prentice Hall, New Jersey, 1997.

[7] J. Cerda, A. W. Westerberg, D. Mason, B. Linnhoff, Minimum utility usage

in heat-exchanger network synthesis. A transportation problem, Chemical Engi-

neering Science 38 (3) (1983) 373 387.

[8] J. Cerda, A. W. Westerburg, Synthesizing heat-exchanger networks having

restricted stream stream matches using transportation problem formulations,

Chemical Engineering Science 38 (10) (1983) 1723 1740.

[9] B. Linnhoff, E. Hindmarsh, The pinch design method for heat-exchanger net-

works, Chemical Engineering Science 38 (5) (1983) 745 763.

215
[10] S. A. Papoulias, I. E. Grossmann, A structural optimization approach in process

synthesis 2. Heat-recovery networks, Computers & Chemical Engineering 7 (6)

(1983) 707 721.

[111 I. C. Kemp, Pinch Analysis and Process Integration, 2nd Edition, Elsevier, 2007.

[12] B. E. Poling, J. M. Prausnitz, J. P. O'Connell, The Properties of Gases and

Liquids, 5th Edition, McGraw Hill, New York, 2001.

[13] L. T. Biegler, Nonlinear Programming: Concepts, Algorithms, and Applications

to Chemical Processes, MPS-SIAM Series on Optimization, SIAM-Society for

Industrial and Applied Mathematics, 2010.

[14] A. Mitsos, P. Lemonidis, P. I. Barton, Global solution of bilevel programs with a

nonconvex inner program, Journal of Global Optimization 42 (4) (2008) 475 513.

[15] M. M. El-Wakil, Powerplant Technology, international Edition, McGraw-Hill,

1985.

[16] J. T. Houghton, Y. Ding, D. J. Griggs, M. Noguer, P. J. van der Linden, X. Dai,

et. al. (Eds.), Climate Change 2001: The Scientific Basis, Cambridge University

Press, 2001.

[17] T. R. Karl, K. E. Trenberth, Modern global climate change, Science 302 (5651)

(2003) 1719 1723.

[18] U. S. Department of Energy., Energy sources: Electric power - fossil fuel, http:

//www.energy.gov/energysources/electricpower.htm (accessed Oct 2010).

[19] M. Gazzino, G. Benelli, Pressurized oxy-coal combustion Rankine-cycle for future

zero emission power plants: Process design and energy analysis, ASME Confer-

ence Proceedings 2008 (43208) (2008) 269 278. doi: 10.1115/ES2008-54268.

http: //link.aip.org/link/abstract/ASMECP/v2008/i43208/p269/s 1

216
[201 J. Hong, G. Chaudhry, J. Brisson, R. Field, M. Gazzino, A. Ghoniem, Analysis of

oxy-fuel combustion power cycle utilizing a pressurized coal combustor, Energy,

the International Journal (2009) 1334 1340.

[21] J. Hong, R. Field, M. Gazzino, A. Ghoniem, Operating pressure dependence

of the pressurized oxy-fuel combustion power cycle, Energy, the International

Journal 35 (2010) 5391 5399.

[22] G. Benelli, D. Cumbo, M. Gazzino, E. Morgani, Pressurized oxy- coal combustion

with flue gas recirculation pilot scale demonstration, Proceedings of the Power-

Gen Europe (2008) paper ID 67.

[23] M. Gazzino, G. Riccio, N. Rossi, G. Benelli, Pressurized oxy-coal combustion

rankine-cycle for future zero emission power plants: Technological issues, Proc.

Energy Sustainability, ASME, San Francisco, CA, USA.

[24] H. Zebian, M. Gazzino, A. Mitsos, Multi-variable optimization of pressurized

oxy-coal combustion, Energy 38 (1) (2012) 37 57.

[25] L. Zheng, R. Pomalis, B. Clements, Technical and economic feasibility study

of a pressurized oxy-fuel approach to carbon capture: PartI technical feasibility

study and comparison of the thermoenergy integrated power system (TIPS) with

a conventional power plant and other carbon capture processes (2007).

[26] C. J. King, Separation Processes, McGraw-Hill, New York, 1980.

[27] H. Zebian, Multi-variable optimization of pressurized oxy-coal combustion, Mas-

ter's thesis, Massachusetts Institute of Technology (2011).

[28] H. Zebian, A. Mitsos, A double-pinch criterion for regenerative Rankine cycles,

Energy 40 (1) (2012) 258 270.

217
[29] H. Zebian, A. Mitsos, A double pinch criterion for optimization of regenerative

rankine cycles, U.S. Patent Application Serial No.: 13/475,816 (May 2012).

[301 G. Benelli, G. Girardi, M. Malavasi, A. Sponaro, Isotherm: A new oxy-

combustion process to match the zero emission challenge in power generation

(2008).

[31] E. Rossetti, M. Malavasi, Method and plant for the treatment of materials, in

particular waste materials and refuse, International Patent WO 2004/094904

(2004).

[32] M. Malavasi, E. Rossetti, High-efficiency combustors with reduced environmen-

tal impact and process for power generation derivable therefrom, International

Patent WO 2005/108867 (2005).

[33] M. Malavasi, G. Di Salvia, Combustion process, International Patent WO

2009/071239 (2009).

[34] F. Star, Background to the design of hrsg systems and implications for CCGT

plant cycling, OMMI (Vol. 2, Issue 1) April 2003.

[35] H. Zebian, A. Mitsos, Pressurized oxycoal combustion: Ideally flexible to uncer-

tainties, Energy 57 (2013) 513-526.

[36] U. S. Energy Information Administration, International energy outlook 2011,

Tech. Rep. DOE/EIA-0484(2011) (September 2011).

[37] The Congress of the United States; Congressional Budget Office. Policy Options

for Reducing CO 2 Emissions (February 2008).

[38] H. Zebian, N. Rossi, M. Gazzino, D. Cumbo, A. Mitsos, Optimal design and

operation of pressurized oxy-coal combustion with a direct contact separation

column, Energy 49 (2013) 268 278.

218
[39] D. Bhattacharyya, R. Turton, S. E. Zitney, Steady-state simulation and op-

timization of an integrated gasification combined cycle power plant with CO 2

capture, Industrial & Engineering Chemistry Research 50 (3) (2011) 1674 1690.

[40] G. Benelli, M. Malavasi, G. Girardi, Innovative oxy-coal combustion process

suitable for future and more efficient zero emission power plants, Proceedings of

the Power-Gen Europe (2007) paper ID 194.

[41] M. Yunt, B. Chachuat, A. Mitsos, P. I. Barton, Designing man-portable power

generation systems for varying power demand, AIChE Journal 54 (5) (2008) 1254

1269.

[42] J. R. Birge, F. Louveaux, Introduction to Stochastic Programming, Springer-

Verlag, 1997.

[43] A. Selot, L. K. Kuok, M. Robinson, T. L. Mason, P. I. Barton, A short-term

operational planning model for natural gas production systems, AIChE J 54 (2)

(2008) 495 515.

[44] B. S. Ahmad, P. I. Barton, Process-wide integration of solvent mixtures, Com-

puters & Chemical Engineering 23 (10) (1999) 1365 1380.

[45] A. Mitsos, I. N. Melas, P. Siminelakis, A. D. Chairakaki, J. Saez-Rodriguez, L. G.

Alexopoulos, Identifying drug effects via pathway alterations using an integer

linear programming optimization formulation on phosphoproteomic data, PLOS

Computational Biology (12) 2009 e1000591 5.

[46] J. Tranier, N. Parrin, D. Dubettier, Air Liquide, Air separtion

units for coal power plants, Carbon Capture Journal (June 2011)-

http://www.carboncapturejournal.com/displaynews.php?NewsID=805.

219
[47] A. Pickard, G. Meinecke, The future role of fossile power generation, Siemens

AG: Energy Sector (2011).

[48] P. Denholm, Y.-H. Wan, M. Hummon, M. Mehos, An analysis of concentrating

solar power with thermal energy storage in a california 33% renewable scenario

(2013).

[49] D. Mallinson, W. Lewis, The part-load performance of various gas-turbine engine

schemes, Proceedings of the Institution of Mechanical Engineers 159 (1) (1948)

198 219.

[50] T. S. Kim, S. T. Ro, The effect of gas turbine coolant modulation on the part

load performance of combined cycle plants. part 1: Gas turbines, Proceedings of

the Institution of Mechanical Engineers, Part A: Journal of Power and Energy

211 (6) (1997) 443 451.

[51] D. L. Chase, Combined-cycle developement evolu-

tion and future, (Oct. 2010), Tech. rep., http://site.ge-


2 (ac-
energy.com/prod-serv/products/tech-docs/en/downloads/ger4 06.pdf;

cessed April 23, 2013).

[52] T. Kim, S. Hwang, Part load performance analysis of recuperated gas turbines

considering engine configuration and operation strategy, Energy 31 (2006) 260

277.

[53] A. C. McMahan, Design and optimization of organic Rankine cycle solar-thermal

power plants, Master's thesis, University of Wisconsin-Madison (2006).

[54] M. Judes, G. Tsatsaronis, Design optimization of power plants by considering

multiple partial load operation points, Proceedings of the Asme International

Mechanical Engineering Congress and Exposition 2007, Vol 6 - Energy Systems:

Analysis, Thermodynamics and Sustainability (2008) 217 225.

220
[55] G. Weber, W. Worek, Sliding pressure analysis using the second law, Heat Re-

covery Systems and CHP 13 (3) (1993) 253 260.

[56] S. Sengupta, A. Datta, S. Duttagupta, Exergy analysis of a coal-based 210 mw

thermal power plant, International Journal of Energy Research 31 (1) (2007)

14 28.

[57] K. Jonshagen, M. Genrup, Improved load control for a steam cycle combined

heat and power plant, Energy 35 (4) (2010) 1694 1700, Demand Response

Resources: The US and International Experience.

[58] A. Stodala, Steam and gas turbines, authorized translation by L. C. Loewenstien,

2nd edition. P. Smith 1945, New York.

[59] Turbine and auxiliaries, nuclear training course 234, module 234-01, the steam

turbine, OPG (1994). https://canteach.candu.org/Content20l3).

[60] B. P. Vitalis, Riley Power Inc., Constant and sliding-pressure options for new

supercritical plants, Power Magazine (Feb 15 2006).

[61] H. Termuehlen, W. Emsperger, Pulverized-Coal-Fired Power Plant Performance

and Operating Flexibility, 1st Edition, ASME Press, 2003.

[62] A. F. Mills, Heat Transfer, Prentice Hall. Inc., 1999.

[63] F. M. White, Fluid Mechanics, McGraw-Hill, 2008.

[64] U. S. Department of Energy, Advanced Manufacturing Office, Deaerator

in industrial steam systems, http: //www1.eere.energy.gov/manufacturing/

tech-assistance/pdf s/steam18-steam-systems.pdf (accessed Sept 2013).

[65] P. C. Tata, Boiler Operation Engineering: Questions and Answers, MPS-SIAM

Series on Optimization, McGraw-Hill Education, 2000.

221
[66] V. Ganapathy, Industrial Boilers and Heat Recovery Steam Generators: Design,

applications, and Calculations, Marcel Dekker Inc., 2003.

[67] H. Zebian, A. Mitsos, Pressurized OCC process ideally flexible to the thermal

load, Energy, (Submitted June 2013).

[68] H. Zebian, N. Seifkar, R. Field, A. Mitsos, Pressurized Isotherm coal oxy-

combustion, Tech. Rep. DOE Award Number: DE-FE000947B (September

2013).

[69] D. Li, J. B. Yang, M. P. Biswal, Quantitative parametric connections between

methods for generating noninferior solutions in multiobjective optimization, Eu-

ropean Journal of Operational Research 117 (1) (1999) 84 99.

222

You might also like